content
stringlengths 1
15.9M
|
---|
\section{Introduction}
\label{intro}
The possibility of extra spatial dimensions~\cite{Kaluza:tu,Klein:tv}, hidden from our current experiments and observations through compactification or warping, has opened up a wealth of options for particle physics model building~\cite{Rubakov:1983bb,Akama:1982jy,Antoniadis:1990ew,Lykken:1996fj,Arkani-Hamed:1998rs,Antoniadis:1998ig,Randall:1999ee,Randall:1999vf,Lykken:1999nb,Arkani-Hamed:1999hk,Antoniadis:1993jp,Dienes:1998vg,Kaloper:2000jb,Cremades:2002dh} and has allowed entirely new approaches for addressing cosmological problems~\cite{Arkani-Hamed:1998nn,Macesanu:2004gf,Starkman:2001xu,Starkman:2000dy,Deffayet:2001xs, Deffayet:2001pu,Dvali:2000hr,Deffayet:2002sp,Deffayet:2000uy,Binetruy:1999hy, Binetruy:2001tc,Binetruy:1999ut,Chung:1999zs,Csaki:1999mp,Cline:2002ht,Nasri:2002rx}.
In many implementations, Standard Model (SM) fields can be confined to a submanifold, or brane, while in others they populate the entire extra-dimensional space. Common to both approaches, however, is the inclusion of bulk fields beyond pure gravity, either because they are demanded by a more complete theory, such as string theory, or because they are necessary to stabilize the extra-dimensional manifold. Thus, a complete understanding of the predictions and allowed phenomenology of extra dimension models necessarily includes a comprehensive consideration of the configurations of these bulk fields, the simplest of which are real scalars. Indeed, 4D Poincare invariance allows for these new bulk fields to acquire nontrivial static configurations along the extra dimensions.
Static one-dimensional scalar configurations with a node (where the field vanishes) are known to localize wave functions of other fields near that node. In the context of extra dimensions, these kink-like scalar backgrounds can be used for example to localize bulk fermions near either boundary~\cite{Arkani-Hamed:1999dc, Georgi:2000wb,Kaplan:2001ga,Grzadkowski:2004mg}, allowing for interesting constructions of flavor models. They can also affect the localization of other scalar or vector fields leading to a field theoretic description of fat branes (see for example the constructions in \cite{Hung:2003cj,Surujon:2005ia,Davies:2007xr}). Kink-like scalar configurations are a particularly interesting case to consider because the boundary conditions make it possible to obtain non-trivial general results regarding both the existence and the stability of such configurations, at least in the case of one flat extra dimension without gravity~\cite{Toharia:2007xe,Toharia:2007xf}. In this paper we build on these previous results and extend them as far as possible to the case with a gravitating (warped) extra dimension. In the presence of gravity the kink-like configuration cannot fix and stabilize the interbrane distance~\cite{Lesgourgues:2003mi}. It is therefore necessary to assume the existence of at least one additional stabilizing field, coupling either directly or gravitationally to the kink field. We will opt for the latter and introduce additional non-interacting scalar fields. At least one of these additional fields must be given a monotonic profile in order to stabilize the size of the extra dimension (i.e., interbrane modulus)~\cite{Goldberger:1999uk}.
The plan of the paper is as follows. In section II we review the results for kink-like backgrounds in a flat extra dimension with no gravity. We then generalize these to include a warped gravitational background, but with no gravitational backreaction from the kink-scalar itself. In section IV we consider the coupled system of multiple scalar fields in the presence of gravity, and as a special case consider a kink field with Dirichlet boundary conditions and another scalar whose purpose is to stabilize the whole configuration. We write the background equations for such a system and show graphically how non-linearities allow a given action to have multiple static solutions.
In the final section we take the first steps towards studying the question of the stability of these static configurations by deriving the complete set of equations for scalar and gravitational perturbations around a given static background. The general procedure is quite complex and involves extended theorems of oscillation theory appropriate to the type of eigenvalue problem we are lead to, namely a matrix Sturm-Liouville problem. We therefore reserve a complete study of the stability of the general system for future work.
\section{Kinked Scalars in Flat Extra Dimensions}
\label{flatxd}
In~\cite{Toharia:2007xe,Toharia:2007xf} a 5D flat scenario including
one real scalar field with an arbitrary scalar potential was
studied and the general conditions for the existence and perturbative
stability of static, nontrivial, background scalar field
configurations were presented. In this section we briefly review the
main results and slightly extend the discussion of the energy densities
of different kink configurations.
Consider a real scalar field in 5 dimensions (labeled by indices $M,N,\ldots =0,1,2,3,5$) with a flat background metric, and defined by the action
\begin{equation}
S = \int d^5x\,\left[
\frac{1}{2} \eta^{MN} (\partial_M \phi) \;\partial_N \phi
- V(\phi)\right]\ .
\label{action_flat_metric}
\end{equation}
The extra dimension is compactified on an $S_1/Z_2$ orbifold with the
scalar field $\phi(x,y)$ being odd under $Z_2$ reflections along the
extra coordinate $x^5\!\equiv\!y$ (i.e. $\phi(x,y)=-\phi(x,-y)$). Here
the orbifold interval is defined as $[0,\pi R]$, with its size $\pi R$
assumed to be fixed. The potential $V(\phi)$ must then be invariant
under the discrete symmetry $\phi\to-\phi$, and is chosen to have at
least two degenerate global minima at $\phi\!=\!\pm v$, with
$v\!\neq\!0$. To simplify notation, we will also choose the potential
to vanish at $\phi\!=\!0$.
\begin{figure}[t!]
\center
\includegraphics[width=16cm,height=8cm]{kinks.eps}\ \
\vspace{0.2cm}
\caption{Profiles in the extra dimension interval $[0,\pi R]$
of different static configurations of the Dirichlet scalar field
$\phi$, defined by the scalar potential
$V(\phi)=-\frac{1}{2}|\mu^2|\phi^2+|\lambda|\phi^4$
($\mu^2\!=\!2M_*^2$, $\lambda\!=\!1M_*^{-1}$,
$\pi R\!=\!\!8.6375M_*^{-1}$). The solutions with nodes in the interval
(dashed curves) are unstable, while the stability of the
nodeless and trivial solutions (solid curves) depend on the parameters
of the model.}
\label{kinks}
\end{figure}
Under these conditions, it was shown in \cite{Toharia:2007xe} that
there will always be static solutions, nontrivial along the extra
coordinate $y$, satisfying the (static) field equation
\begin{equation}
\phi'' - \frac{\partial V}{\partial \phi} = 0 \ ,
\label{scal_el_general}
\end{equation}
where a prime denotes a derivative with respect to $y$. The profiles
of these solutions, satisfying Dirichlet boundary conditions, resemble
that of a kink solution patched to an anti-kink in the middle of the
interval. The possible solutions were classified in two groups, namely
those with nodes in the interval (multiple kink-antikink solutions patched
together) and those with no nodes, vanishing only at the end-points of
the orbifold (see Fig.~\ref{kinks}). It was shown that all static
kink solutions with nodes are perturbatively unstable, whereas the
stability of nodeless solutions depends on the parameters of the model
in a particularly simple way.
\begin{figure}[t!]
\center
\includegraphics[width=16cm,height=8cm]{nodelesskinks.eps}
\vspace{0.2cm}
\caption{Nodeless static configurations of the kink scalar field
$\phi$, defined by the scalar potential
$V(\phi)=-\frac{1}{2}|\mu^2|\phi^2+|\lambda|\phi^4$
($\mu^2\!=\!2M_*^2$, $\lambda\!=\!1M_*^{-1}$). Configurations with
different amplitudes are solutions to different physical problems,
corresponding to different stabilization radii of the extra dimension.
The vertical dashed line indicates the minimal radius $R_c$, below which nodeless
solutions do not exist with this potential.}
\label{nodelesskinks}
\end{figure}
The Dirichlet solutions of Eq.~(\ref{scal_el_general}) with no nodes
in the interval form a continuous one-parameter family of functions. A
simple choice for the parameter is the amplitude $A$ of the solution,
i.e., the maximum value of the nontrivial solution
$\phi_A(y)$. Solutions with different amplitudes $A$ generally vanish
at different points along the extra dimension, which correspond to
different possible orbifold radii $R$ (see
Fig.~\ref{nodelesskinks}). However, in order to obtain the stability
condition for these solutions it is extremely useful to consider the
full family of solutions.
The value of the 4D effective energy density of a
given static solution $\phi_A(y)$ is
\begin{equation}
E(A) = \int_0^{T(A)} \left(\frac{1}{2}{\phi'_A}^2 +
V(\phi_A)\right) dy \ ,
\end{equation}
where $T(A)$ is the length of the solution in the extra
dimension. This can be conveniently rewritten as an integral over
$\phi$ using properties of Eq.~(\ref{scal_el_general}) and its
solutions $\phi_A(y)$
\begin{equation}
E(A) = 2\sqrt{2}\int_0^A \frac{V(A)-2\ V(\phi)}{\sqrt{V(\phi)-V(A)}}\,d\phi \ .
\label{energyvsA}
\end{equation}
We are now equipped to state the general results of
\cite{Toharia:2007xe,Toharia:2007xf} in a slightly modified, although
more revealing, version:
\begin{prop}
A static solution to equation~(\ref{scal_el_general}),
with $\delta >0$ nodes inside the orbifold interval is always unstable.
\end{prop}
\begin{prop}
A static, nodeless solution $\phi_{A_*}(y)$ to
equation~(\ref{scal_el_general}), with amplitude $A_*$, and associated
energy density $E(A_*)$ is stable if
\begin{equation}
\left.\frac{dE}{dA}\right|_{A=A_*} < 0 \ .
\label{Estabilitycriterion}
\end{equation}
\end{prop}
\noindent This is a powerful result since it means that given any
scalar potential $V(\phi)$ we immediately know
which of the nontrivial nodeless solutions $\phi_A$ will be stable or
unstable, without the need to actually know explicitly their analytic
form.
With this result it is possible to understand the vacuum structure of
any single scalar field theory with Dirichlet boundary conditions in 5D when the metric along
the extra dimension is flat. Possible static solutions consist of the
trivial solution $\langle\phi\rangle=0$ (which may or may not be stable), kink-like
solutions with nodes in the interval (which are always unstable), and
kink-like solutions without nodes in the interval (some stable and some
unstable, depending on condition (\ref{Estabilitycriterion})). As
remarked in \cite{Toharia:2007xe,Toharia:2007xf}, the trivial solution
may be the true vacuum solution even in the case of a negative mass
term $-|\mu^2| \phi^2$ in the 5D potential, as long as the inequality
$|\mu^2|<|1/R^2|$ is preserved. Therefore, for a given orbifold radius $R$, many different perturbatively stable
vacuum solutions are possible, and it is necessary to identify which
one is the true vacuum of the theory.
The true vacuum of the theory will depend on the size of the radius $R$. This
can be seen as follows: Without loss of generality, one may define the
energy density of the trivial solution to be zero by choosing the 5D
potential $V(\phi)$ to vanish at $\phi=0$. It was shown
in~\cite{Toharia:2007xe,Toharia:2007xf} that there is a critical
radius $R_{c}$ below which nontrivial nodeless solutions do not exist
(see Fig.~\ref{nodelesskinks}). The energy density associated with the
critical nontrivial nodeless solution will
be either positive or exactly zero, so that the transition from one
vacuum to another can be either second order or first order, as
one varies the radius $R$.
\begin{figure}[ht!]
\center
\includegraphics[width=16cm,height=6cm]{kinks+EofR.eps}
\vspace{0.2cm}
\caption{Profiles (left panel) in the extra dimension interval
$[0,\pi R]$ of the two possible stable static
configurations of the Dirichlet scalar field $\phi$, defined by the scalar
potential
$V(\phi)=-\frac{1}{2}|\mu^2|\phi^2-\frac{1}{4}|\lambda|\phi^4+\frac{1}{6}|\xi|\phi^6$
(with $\mu^2=4M_*^2$, $\lambda=4M_*^{-1}$, $\xi=0.6M_*^{-4}$, and
$\pi R=1.368M_*^{-1}$). In the right panel, we show the energy of
the two stable solutions as a function of $\pi R$, and it is seen how
the absolute stability of coexisting static configurations is
determined by the size of the radius $R$. The dots show critical
points where the scalar perturbations contain a massless mode.}
\label{phasediagram}
\end{figure}
In Fig.~\ref{phasediagram} we show an example of a simple setup
defined by the scalar field potential
$V(\phi)=-\frac{1}{2}|\mu^2|\phi^2-\frac{1}{4}|\lambda|\phi^4+\frac{1}{6}|\xi|\phi^6$,
with $\mu^2=4M_*^2$, $\lambda=4M_*^{-1}$ and $\xi=0.6M_*^{-4}$. In the right panel,
the energy density of two static solutions is plotted as a function of
$R$, showing clearly that below a critical radius $R_{1}$ only the
trivial solution is possible and above a critical radius $R_{2}$ only
the kink solution is possible. For $R_{1}<R<R_{2}$, both solutions are perturbatively stable. At the radius $R_{*}$ the two solutions are
degenerate, marking the transition from one true vacuum to another
($\phi_{\mathrm{triv}}$ for $R<R_{*}$ and $\phi_{\mathrm{kink}}$ for
$R>R_{*}$). From this we see that the inverse length scale $1/R$ plays
the role of an order parameter of a phase transition, much like
temperature $T$ in finite temperature field theory. For a very small
radius $R$ (analogous to high $T$) the system is stable only around
its trivial solution, with all symmetries restored. As the radius
increases (analogous to $T$ decreasing) the system can undergo a phase
transition, which could be of either first or second order. The analogy with
temperature, however, is not meant to be taken literally. For whereas
the temperature in any 4D effective cosmology must be monotonically
decreasing for most of its history, the orbifold radius $R$ could in
principle increase, decrease or oscillate on very long time scales,
depending on the dynamics of the stabilization mechanism (which we have so far
ignored).
\section{Kinks on a Warped Background}
\label{warpedxd_fixed}
We now extend previous investigations to the case of a scalar field in
a warped extra dimension, while neglecting any backreaction on the warping from
the scalar field itself. In this case one includes the
effects of the curved metric along the extra dimension on the scalar
field solutions while still ignoring the dynamics of the gravitational
sector. We therefore consider the action
\begin{equation}
S = \int d^5x\sqrt{-g}\,
\left[\frac{1}{2}g^{MN}(\partial_M \phi)\,\partial_N \phi
- V(\phi)\right] \ ,
\label{action_warped_metric}
\end{equation}
where the form of the metric is now taken to be
\begin{equation}
ds^2 = e^{-2\sigma(y)}\gamma_{\mu\nu}(x)dx^\mu dx^\nu - dy^2\ ,
\label{warped_metric}
\end{equation}
and where $\sigma(y)$ is the
warp-factor and $\gamma_{\mu\nu}$ the 4D metric on slices of constant
$y$. The purpose of considering scalar field configurations on a fixed
background is to explore whether our previous
results continue to hold in the presence of a warped background in a
regime where we still have semi-analytical control over the
solutions. We postpone a discussion of the full dynamical problem,
including the backreaction on the metric due to the presence of the scalar
field, until the next section.
\subsection{Kink Scalar in an $AdS_5$ Background}
In the original Randall-Sundrum (RS) model~\cite{Randall:1999ee}, the metric takes
the form (\ref{warped_metric}) with $\sigma(y)=k|y|$ and
$\gamma_{\mu\nu}=\eta_{\mu\nu}$, where $k$ has dimensions of mass and is related to the 5D cosmological
constant of $AdS_5$. In this background any static nontrivial field
configurations $\bphi(y)$ are solutions of
\begin{eqnarray}
\bphi'' - 4 k\bphi'
- \left.\frac{\partial V}{\partial\phi}\right|_{\bphi} = 0 \ .
\label{kinkeqads}
\end{eqnarray}
Scalar perturbations around this kink background, $\pphi(x,y)=\phi(x,y)-\bphi(y)$, can be decomposed as
\begin{equation}
\pphi(x,y) = \displaystyle\sum_n\pphi_{x}^{(n)}(x)\pphi_{y}^{(n)}(y)
\label{pphidecomp}
\end{equation}
such that the normal modes $\pphi_{x}^{(n)}(x)$ and $\pphi_{y}^{(n)}(y)$ are solutions of
\begin{eqnarray}
{}^{(4)}\Box \pphi_{x} + m_{n}^2\pphi_{x} &=& 0 \\
\pphi_{y}'' - 4k\pphi_{y}' - (\mu^2(y) - m_{n}^2 e^{2ky})\pphi_{y} &=& 0,
\label{pphieq}
\end{eqnarray}
where $\mu^2\equiv\left.\frac{\partial^2
V}{\partial\phi^2}\right|_{\bphi}$ and
${}^{(4)}\Box\equiv\eta^{\mu\nu}\partial_{\mu}\partial_{\nu}$. Taking the
derivative of the kink equation (\ref{kinkeqads}) gives
\begin{equation}
\pphi_{M}'' - 4k\pphi_{M}' - \mu^2(y)\pphi_{M} = 0 \ ,
\end{equation}
where we have defined $\pphi_{M}\equiv\bphi'$. Thus $\pphi_{M}$ is a massless
solution ($m_{n}^2=0$) of the perturbation equation~(\ref{pphieq}),
although it satisfies mixed boundary conditions rather than the
Dirichlet boundary conditions imposed on $\pphi_{y}$.
At this point we are already able to state a new result of this work, which is an
extension of the previous result related to the impossibility of
having stable kink solutions with nodes inside the interval.
Suppose that $\bphi(y)$ happens to have $\delta$ nodes inside the
interval. We have just shown that $\bphi'\equiv \pphi_{M}$ will solve
the equation for a massless excitation, but with mixed boundary
conditions. Since $\bphi$ has $\delta$ nodes, $\bphi'=\pphi_{M}$
must have $\delta+1$ nodes inside the interval. The following inequalities relating the eigenvalues $\lambda^D_{n}$
for the Dirichlet case and the eigenvalues $\lambda^M_{n}$ for a general
mixed boundary condition case~\cite{zettl} hold from Sturm-Liouville theory
\begin{equation}
\lambda^D_{n} \le \lambda^M_{n+2} \le \lambda^D_{n+2} \ .
\end{equation}
Since we have $\lambda^M_{\delta+1}=0$ (i.e. the eigenvalue of the
solution with $\delta+1$ nodes), we can immediately deduce that the
mass-squared of the lowest excitation of the Dirichlet problem must be
negative since $\lambda^D_{\delta-1} \le \lambda^M_{\delta+1}=0$ with
$\delta \ge 1$.
\begin{prop} In a warped background on a slice of
$AdS_5$, any static solution to equation~(\ref{kinkeqads}),
with $\delta>0$ nodes inside the interval is always unstable.
\end{prop}
\noindent However, for nodeless static solutions
(when $\delta=0$) the results for the flat case obtained in
\cite{Toharia:2007xe,Toharia:2007xf} cannot
be extended here. Lacking a general stability condition, we will instead propose a weaker sufficient
stability condition for these and other more generic solutions in the next subsection.
\subsection{Kink Scalar on a General Background}
In a general warped background with metric ansatz~(\ref{warped_metric}) the equation for a static scalar
background configuration is
\begin{equation}
\bphi'' - 4 \sigma'\bphi'
- \left.\frac{\partial V}{\partial\phi}\right|_{\bphi} = 0 \ .
\label{kinkeqsig}
\end{equation}
In this situation, although we have been unable to extend the stability theorems
found earlier, we are still able to find a general
{\it sufficient}
condition for perturbative stability of the background configurations.
Small perturbations around the background $\bphi(y)$ may be defined as
in (\ref{pphidecomp}). The spectrum of these perturbations consists of
solutions to the eigenvalue problem
\begin{eqnarray}
\pphi_{y}(y_{1}) = \pphi_{y}(y_{2}) = 0 \\
\pphi_{y}'' - 4\sigma'\pphi_{y}'
- \left[\mu^2(y) - m^2_{n}e^{2\sigma(y)} \right]\pphi_{y} = 0 \ .
\label{pphieqsig}
\end{eqnarray}
A useful form of this equation is obtained by performing a change of
variables $\ e^{\sigma(y)} dy=dz$ and defining $\sigma(z)=-\frac{2}{3}\ln{(J(z))}$ and
$W(z)=\mu^2(z)e^{-2\sigma(y(z))}$ to yield
\begin{equation}
\frac{(J\pphi)''}{J\pphi}- \frac{J''}{J} - \left(W(z)-m_{n}^2 \right) = 0 \ .
\label{Jequation}
\end{equation}
To proceed, we make use of the following integral
inequality~\cite{math}. For any function $f(z)$, such that
$f(a)=f(b)=0$, and with $n$ nodes within the interval $[a,b]$, there exists $\rho\in{\cal R}$ such that
\begin{equation}
\int_a^b e^{-\rho f''(z)/f(z)}\ dz \ge (n+1) e \sqrt{\rho \pi} \ .
\end{equation}
Applied to~(\ref{Jequation}), this implies
\begin{equation}
e^{\rho m_{n}^2} \int_a^b e^{-\rho [\frac{J''}{J} + W(z) ]}\ dz \ge (n+1) e \sqrt{\rho \pi} \ ,
\end{equation}
the logarithm of which yields
\begin{equation}
m_{n}^2 \ge \frac{1}{\rho}\ln[ (n+1) e \sqrt{\rho \pi}]
- \frac{1}{\rho}\ln\left( \int_a^b e^{-\rho [\frac{J''}{J} + W(z) ]}\ dz \right)
\end{equation}
which is a lower bound for the eigenvalues in terms of the
background quantities $\sigma(z)$ and $\mu(z)$ (which are contained in $J$ and
$W$). In the case of the lowest eigenvalue we have
\begin{equation}
m_{0}^2 \ge \frac{1}{\rho}\ln( e \sqrt{\rho \pi})
- \frac{1}{\rho}\ln\left( \int_a^b e^{-\rho [\frac{J''}{J} + W(z)]}\ dz \right)
\end{equation}
and so a sufficient condition for perturbative stability ($m_{0}^2\ge 0$) is
\begin{equation}
\int_a^b e^{-\rho [\frac{J''}{J} + W(z) ]}\ dz \le e \sqrt{\rho \pi} \ .
\end{equation}
We may formulate this explicitly in terms of the warp factor
$\sigma(z)$ so that finally, a static solution $\bphi(z)$ of~(\ref{kinkeqsig}), obeying
Dirichlet boundary conditions, is stable if
\begin{equation}
\int_a^b e^{-\rho[-\frac{3}{2}\sigma'' + \frac{9}{4}\sigma'^2 +\mu^2(z)e^{-2\sigma}]}\ dz
\le e \sqrt{\rho \pi} \ ,
\label{suffcond}
\end{equation}
where $\left.\mu^2(z)\equiv\frac{\partial^2 V}{\partial
\phi^2}\right|_{\bphi_(z)}$ and where the actual value of $\rho$ is
that which extremizes the right hand side. This is a sufficient
condition, but not a necessary one. In order to demonstrate how
effective this {\it weaker} stability condition can be, we now turn
to a simple example in which the condition can actually be evaluated.
Consider a flat metric, where $\sigma(y)\equiv 0$, and a trivial
background scalar configuration, i.e., $\bphi(y)\equiv 0$, but where
the 5D scalar potential is allowed to have a tachyonic mass. In this
case equation (\ref{pphieqsig}) becomes
\begin{equation}
\pphi_{y}'' - \left(\mu^2-m_{n}^2 \right) \pphi_{y}=0 \ .
\end{equation}
If $\pphi_{y}$ has Dirichlet boundary conditions, the solutions to this problem are
\begin{equation}
\pphi_{y}= \sin(\sqrt{m_{n}^2-\mu^2}\ y)
\end{equation}
where $m_{n}^2-\mu^2=n^2\pi^2/L^2$ and $L=b-a$ is the size of the extra
dimension. The mass of the lightest mode is
$m_{0}^2=\mu^2+\pi^2/L^2 $ and so the condition for stability is
$\ m_{0}\le 0\ $, which means that the bulk scalar mass $\mu^2$ can be
negative, but not arbitrarily so:
\begin{equation}
\mu^2\ge -\pi^2/L^2\ .
\label{flatexact}
\end{equation}
Therefore in this case (where $\sigma'=\sigma''=0$), our {\it sufficient} condition~(\ref{suffcond}) becomes
\begin{equation}
e^{-\rho \mu^2} \int_a^b dz \le e \sqrt{\rho \pi} \ ,
\end{equation}
which leads to
\begin{equation}
\mu^2 \ge \frac{1}{\rho}\ln(\frac{L}{e\sqrt{\rho \pi}}) \ .
\end{equation}
The value of $\rho$ that extremizes this bound is $\rho=\pi L^2/e$, and so our weaker bound is
\begin{equation}
\mu^2 \ge - \frac{e\pi}{2} \frac{1}{L^2} \ .
\end{equation}
This result is a factor of $2\pi/e$ weaker than the exact bound~(\ref{flatexact}).
Nevertheless this result is nontrivial as it clearly demonstrates that it is possible to
have negative bulk masses and retain a stable system.\footnote{The
stability conditions of the trivial vacuum in the presence of
negative bulk mass terms in an extra-dimensional scalar field theory
have been analyzed and generalized to general warped backgrounds in
\cite{Toharia:2008ug}.}
\section{Kinks in Gravitating Warped Extra Dimensions}
\label{warpedxd_dynamic}
So far we have examined static scalar field configurations in a fixed
background. We have found that some of the results that were shown to
hold in a flat extra-dimensional background continue to hold in a
fixed warped background, and we have found useful generalizations of other results.
We now want to include the dynamics of the
gravitational sector and explore how these results can be extended
when the gravitational backreaction is included. Therefore we now seek nontrivial
static field configurations in which the warp factor has its own dynamics
determined by the 5D Einstein equations.
As soon as we include a dynamical gravitational sector, we are required to
worry about stabilization of the extra dimension. In the above
discussion we assumed that the extra dimension was stabilized and that the
dynamics of the stabilization mechanism were frozen
out. Here we want to include the backreaction of any matter fields on
the 5D metric, and so we must include the dynamics of stabilization. A natural
question to ask is whether the kink fields of interest
could provide a stabilization mechanism. Unfortunately, in
\cite{Lesgourgues:2003mi} it was shown that when one considers static
solutions for both the warp factor and a single scalar field, the
lightest scalar perturbative mode (the radion) will be tachyonic
whenever the derivative of the scalar profile vanishes inside the
interval. In other words, the system is unstable whenever the scalar
field profile passes through an extremum in the bulk. This means that if we insist
on obtaining a nontrivial configuration for a single scalar field with
Dirichlet boundary conditions, we are guaranteed to obtain a tachyonic
radion and the extra dimension will be unstable. To address this issue
we will add extra scalar fields whose purpose will be to stabilize the
radion as in~\cite{Goldberger:1999uk}.
The resulting system becomes considerably more difficult to analyze
than the case with only one bulk scalar field, particularly with regard
to questions about stability. On the other hand, the case with three
or more scalar fields is formally no more difficult to analyze than
the case with only two scalar fields. Hence we will keep our treatment
general to include an arbitrary number of scalar fields $\chi_{a}$
($a,b=1,\ldots,{\cal N}$), although when we consider particular
examples below, we will specialize to the case with only two scalar
fields (a kink field and a non-kink field). For simplicity we will assume throughout that the scalar
fields are only coupled gravitationally.
We therefore consider the 5D action for gravity and ${\cal N}$ free scalar fields
\begin{eqnarray}
S &=& -\frac{M_*^3}{2}\int d^5x\sqrt{-g}\,\left[{\cal R} - 2\Lambda\right] \nonumber
\\
&+&
\int d^5x\sqrt{-g}\,
\left[\displaystyle\sum_{a=1}^{{\cal N}}\frac{1}{2}\,g^{MN}(\partial_M\chi_{a})(\partial_N\chi_{a})
- W(\chi_{a}) -
\displaystyle\sum_{i=1,2}\lambda_{i}(\chi_{a})\delta(y-y_{i})\right] \ ,
\end{eqnarray}
where $M_*\equiv(8\pi G)^{-1/3}$, $G$ is the 5D Newton's constant,
${\cal R}$ is the 5D Ricci scalar, and $\Lambda$ is the 5D cosmological
constant. The full scalar potential in the bulk is $W(\chi_{a})$, and
the brane potentials are $\lambda_{i}(\chi_{a})$. As before, we take
the 5D line element of the form
\begin{equation}
ds^2 = e^{-2\sigma(y)}\gamma_{\mu\nu}(x) dx^\mu dx^\nu - dy^2\ ,
\end{equation}
where $\gamma_{\mu\nu}$ is the induced metric on the 4D hypersurfaces of
constant $y$, which foliate the extra dimension. The 5D
Einstein and field equations are
\begin{eqnarray}
\sigma'' - \sigma'^2 + \frac{\Lambda}{6}
&=& \frac{1}{2M_*^3}\left( \displaystyle\sum_{a}^{{\cal N}} \frac{1}{2}\chi_{a}'^2
+ \frac{1}{3}\,W(\chi_{a})
+ \frac{2}{3}\displaystyle\sum_{i=1,2}\lambda_{i}(\chi_{a})\delta(y-y_{i})\right)
\label{bkg1}
\\
\sigma'^2 &-& \frac{\Lambda}{6} + \frac{{}^{(4)}{\cal R}\,\,}{12}e^{2\sigma}
= \frac{1}{6M_*^3} \left( \displaystyle\sum_{a}^{{\cal N}} \frac{1}{2}\chi_{a}'^2
- W(\chi_{a})\right)
\label{bkg2}
\\
\chi_{a}'' &-& 4\sigma'\chi_{a}' - \frac{\partial W}{\partial\chi_{a}}
- \displaystyle\sum_{i=1,2}\frac{\partial\lambda_{i}}{\partial\chi_{a}}\delta(y-y_{i})
= 0 \ ,
\label{bkg3}
\end{eqnarray}
where ${}^{(4)}{\cal R}$ is the 4D Ricci scalar associated with the induced
4D metric $\gamma_{\mu\nu}$, which we have left arbitrary. The boundary
conditions for the system are determined by Israel junction
conditions at each brane. These are obtained by integrating the
equations of motion over an infinitesimally small interval across each brane, giving
\begin{eqnarray}
\left[\sigma'\right]_{y_{i}} &\equiv&
\displaystyle\lim_{\epsilon\to\,0} \left[\sigma'(y_{i}+\epsilon) - \sigma'(y_{i}-\epsilon)\right]
= \frac{1}{3M_*^3}
\left.\lambda_{i}(\chi_{a})\right|_{y_{i}}
\label{jxn1}
\\
\left[\chi'_{a}\right]_{y_{i}} &\equiv&
\displaystyle\lim_{\epsilon\to\,0} \left[\chi_{a}'(y_{i}+\epsilon) - \chi_{a}'(y_{i}-\epsilon)\right]
= \left.\frac{\partial\lambda_{i}}{\partial\chi_{a}}\right|_{y_{i}}\ .
\label{jxn2}
\end{eqnarray}
These yield ${\cal N}$ conditions on each brane, which is exactly the number
of data that need to be specified in order for equations (\ref{bkg1})
and (\ref{bkg3}) to form a well-posed problem.
Note that the above boundary value problem consists of a system of coupled
nonlinear differential equations. Finding solutions analytically for
such a setup is highly unlikely, although it is still possible to
proceed in the opposite direction, i.e. given a particular analytical
solution one can obtain the setup from which it originates. To do so,
one relies on the powerful method of the
superpotential~\cite{Skenderis:1999mm,DeWolfe:1999cp,Csaki:2000zn}, which can be useful even for two
or more scalar fields (see, for example,~\cite{Batell:2008zm} in the context
of soft-wall models). However, even if one solution is constructed in this
way, there is no guarantee that this is the only solution with the same action. We will now describe how to look for all
possible solutions of a given action using a combination of numerical and graphical
techniques.
\subsection{Multiple Solutions}
Whenever there is more than one static solution to the above
boundary value problem {\it with the same action}, we say that multiple solutions exist. In general, the bulk scalar fields can have Dirichlet boundary
conditions, Neumann boundary conditions or more general mixed boundary
conditions. Here we focus on the case where we have one
kink field $\phi$ (obeying Dirichlet boundary conditions), with the remaining ${\cal N}-1$ fields $\chi_{a}$ having
Neumann or mixed boundary conditions. When the profiles of these extra fields
are monotonic, they will tend to stabilize the extra dimension, whereas
if their profiles have vanishing derivatives inside the interval, they will tend to destabilize the
extra dimension~\cite{Lesgourgues:2003mi}. Despite this subtlety, we
will generically refer to the non-kink fields as ``stabilization''
fields.
To find solutions we proceed as follows: we specify the Lagrangian in
the bulk and on one of the branes, and we numerically solve an initial
value problem to determine the profiles of the fields along the extra
dimension. Dirichlet boundary conditions are imposed on the kink field
$\phi$ at the initial brane by demanding that it vanish there. For
this to hold, we assume the kink field has a sufficiently heavy brane
mass so that it decouples from the stabilization fields on the
branes. As a result, the kink field disappears from the junction
conditions (\ref{jxn1})-(\ref{jxn2}), which then yield only ${\cal N}$
conditions on the initial brane. This leaves ${\cal N}+1$ initial conditions
that need to be specified, which we take to be the boundary values for
the derivatives $\phi'$, $\chi_{1}'\ldots\chi_{{\cal N}-1}'$, and
$\sigma'$. After solving the initial value problem for a given choice
of initial conditions, we impose Dirichlet boundary conditions on
$\phi$ at the final boundary by locating the second brane at a point
where the profile of $\phi$ vanishes. In general, the profile will
vanish at several points along the extra dimension, and one may study
kinks with the desired number of nodes by choosing the location of the
second brane accordingly. Here, as in the flat case, we are primarily
interested in nodeless kink solutions, and we therefore place the second brane
at the first zero of the profile function.
We now have a solution to a {\it boundary} value problem whose
boundary conditions on the second brane are not yet known. We
parameterize the brane potential on the second brane
$\lambda_{2}(\chi_{a})$ in terms of $P$ parameters $\alpha_{b}$ (for example,
the brane tension $\Sigma_2$, the brane mass term $m_{a}^2$ of each scalar, the quartic
coupling of each scalar, etc.)
\begin{eqnarray}
&&\lambda_{2}(\chi_{a})
= f(\Sigma_2,m_{1}^2,m_{2}^2,...,m_{{\cal N}}^2,...) \ .
\label{secondbrane}
\end{eqnarray}
Then the junction conditions (\ref{jxn1})-(\ref{jxn2}) at the second
brane ($i=2$) give ${\cal N}$ linear equations for the $P$ unknowns
$\alpha_{b}$. By evaluating the fields on the second brane, and
using the parameterization in (\ref{secondbrane}), we then
invert the ${\cal N}$ junction conditions to determine the $\alpha_{b}$.
If this is possible, then the solution to our initial value problem is
also a solution to a corresponding boundary value problem. From this
we see that we must have $P\geq{\cal N}$ in order to guarantee that the
field configuration we obtained is the solution to a corresponding
boundary value problem. If $P={\cal N}$, the $\alpha_{b}$ are uniquely
determined, and there is a unique Lagrangian for which the above field
configuration is a solution. On the other hand if $P>{\cal N}$, some of the
$\alpha_{b}$ are arbitrary and so there is a family of solutions for
these final-boundary conditions. In that case there is a family of
Lagrangians which yield the obtained field configuration, and one can
proceed by focusing on one member of this family. If $P<{\cal N}$, the
linear system of parameters $\alpha_{b}$ may be overdetermined, in
which case the obtained field configuration is not a solution to any
corresponding boundary value problem.
We can find additional solutions by changing the initial-boundary conditions
and repeating the above process. Note that by freely varying the field
derivatives ($\phi'$, $\chi_{1}'\ldots\chi_{{\cal N}-1}'$) at the initial
brane and determining the remaining quantities from the junction
conditions, it is possible to leave the initial-brane potential unchanged. This
is necessary in order that the action remains unchanged (it is not
sufficient because part of the action is determined by the final-brane
potential). A solution and the resulting final-boundary conditions
(the $\alpha_{b}$) are then found as before. Since each set of
initial shooting values yields a set of $\alpha_{b}$, each $\alpha_{b}$ is a function of the ${\cal N}$ initial-boundary
derivatives. Each $\alpha_{b}$ therefore defines an ${\cal N}$-dimensional
surface whose level-surfaces can be projected onto the
$\phi'(y_{1})$-$\chi_{a}'(y_{1})$ parameter space (which is an
${\cal N}$-dimensional space). In the above construction there are $P$ such
quantities, and so $P$ level-surfaces intersect at every point in this
parameter space, representing one solution for this action. The
question of whether multiple solutions exist for the same action is
equivalent to the question of whether the same $P$ surfaces
simultaneously intersect at more than one point in the parameter
space.
We will now show how this works in two simple examples. In both
cases, we will consider a kink field $\phi$ in addition to just one
stabilization field $\chi$, with no interaction terms among them in the scalar potential. In
both examples there will be regions of parameter space in which two
distinct static configurations are possible for the same action.
\subsection{Example 1: Quartic Potential}
In both of the following examples we consider a Lagrangian for two scalar fields
\begin{eqnarray}
{\cal L}_{matter} &=& \frac{1}{2}\,g^{MN}(\partial_M\phi)\partial_N\phi
- V(\phi) - \displaystyle\sum_{i=1,2}\beta_{i}(\phi)\delta(y-y_{i}) \nonumber
\\
&& \mbox{} + \frac{1}{2}\,g^{MN}(\partial_M\chi)\partial_N\chi
- U(\chi) - \displaystyle\sum_{i=1,2}\lambda_{i}(\chi)\delta(y-y_{i}) \ ,
\end{eqnarray}
where $\phi$ is the kink field and $\chi$ is the stabilization field with potentials
\begin{eqnarray}
U(\chi) &=& \frac{1}{2}\,m_{\chi}^2\chi^2
\label{gw_bulk_pot}
\\
\lambda_{i}(\chi) &=& M_*^{-1}\left(\frac{1}{2}\mu_{i}^2\chi^2 + \Sigma_{i}\right)\ .
\label{gw_brane_pot}
\end{eqnarray}
The fact that the second brane potential for $\chi$ is parameterized
in terms of two parameters, $\mu_{2}^2$ and $\Sigma_{2}$, will allow us
to find unique solutions to the boundary conditions on the second
brane. The junction conditions (\ref{jxn1})-(\ref{jxn2}) become
\begin{eqnarray}
\sigma'(y_{i})
&=& (-1)^{i-1}\frac{1}{6M_*^4}
\left(\frac{1}{2}\mu_{i}^2\chi^2(y_{i}) + \Sigma_{i}\right)
\label{example_jxn1}
\\
\chi'(y_{i})
&=& (-1)^{i-1}\frac{1}{2}\mu_{i}^2\chi(y_{i})\ .
\label{example_jxn2}
\end{eqnarray}
On the second brane ($i=2$) these can be inverted to give
\begin{eqnarray}
\mu_{2}^2 &=& -2\frac{\chi'(y_{2})}{\chi(y_{2})}
\label{mu}
\\
\Sigma_{2} &=& -6M_*^4\sigma(y_{2}) + \chi'(y_{2})\chi(y_{2})
\label{Sigma}
\end{eqnarray}
so that once we determine the fields on the second brane, we can
extract the boundary conditions (and therefore Lagrangian) to which
those fields are a solution.
The only things left to specify are the bulk potential for the kink
field and the initial-boundary conditions. In this first example, we
take the kink potential to be
\begin{equation}
V(\phi) = -\frac{1}{2}\,m_{\phi}^2\phi^2 + \frac{1}{4}\lambda\phi^4 \ .
\label{kink_pot1}
\end{equation}
Taking the initial brane to be located at $y=0$, we find solutions to
the initial-boundary value problem at this brane with Dirichlet
boundary conditions imposed on the field $\phi$. Examples of nodeless
solutions to the initial value problem are shown in
Fig.~\ref{bsoln1}. To ensure that $\phi$ obeys Dirichlet boundary
conditions on the second brane, we locate the second brane at the
first point (other than $y=0$) where the profile of $\phi$ vanishes
(the vertical dashed lines in Fig.~\ref{bsoln1}).~\footnote{For certain initial
conditions, the profile of $\phi$ will blow up before it vanishes
for a second time. When this happens the initial conditions used do
not lead to a solution of our boundary value problem.} Once the
position of the second brane is identified, the final-boundary
conditions are determined from (\ref{mu}) and (\ref{Sigma}). By
varying the initial shooting conditions, $\phi_{1}'\equiv\phi'(y_{1})$ and $\chi_{1}'\equiv\chi'(y_{1})$,
and repeating this process of finding solutions, identifying the
location of the second brane, and determining the final-boundary
conditions, we generate level-curves of $\mu_{2}^2$ and
$\Sigma_{2}$. These are plotted in Fig.~\ref{contour1}. Notice that most
$\mu_{2}^2$ contours cross each $\Sigma_{2}$ just once, signifying
that there is a single solution for the corresponding action with the kink potential of
Eq.~(\ref{kink_pot1}). However, some contours cross each other more
than once (see, for example, the circles in Fig.~\ref{contour1}.) Furthermore there is only a
finite region in the
$\phi_{1}'$-$\chi_{1}'$ parameter space where solutions exist. If either
$|\phi_{1}'|$ or $|\chi_{1}'|$ are increased sufficiently, the solution to the
initial value problem blows up.
In that case the {\it boundary} value problem has no solution, since a second boundary
where $\phi=0$ does not exist. It is therefore possible to scan
the entire allowed $\phi_{1}'$-$\chi_{1}'$ space and examine whether multiple
solutions with the same action exist.
\begin{figure}[t!]
\includegraphics[width=16cm,height=12cm]{bsoln1.eps}
\vspace{0.2cm}
\caption{Profiles of the scalar backgrounds $\phi(y)$ and $\chi(y)$
as well as the warp factor $\sigma(y)$, showing the two possible solutions
(panels A and B) to the same boundary value problem defined by the physical parameters
$m_{\chi}^2=-0.5M_*^2$, $\mu_{1}^2=-0.25M_*^2$,
$\mu_{2}^2=-8M_*^2$, $\Sigma_{1}=-2M_*^4$,
$\Sigma_{2}=0.52M_*^4$, $m_{\phi}^2=0.5M_*^2$, $\lambda=2M_*^{-1}$,
and $\Lambda=0$.
}
\label{bsoln1}
\end{figure}
\begin{figure}[t!]
\includegraphics[width=16cm,height=12cm]{contour1.eps}
\vspace{0.2cm}
\caption{Level-curves of $\mu_{2}^2$ and $\Sigma_{2}$ in the
$\phi'_{1}$-$\chi'_{1}$ parameter space for example 1 with
$m_{\chi}^2=-0.5M_*^2$, $\mu_{1}^2=-0.25M_*^2$,
$\Sigma_{1}=-2M_*^4$, $m_{\phi}^2=0.5M_*^2$,
$\lambda=2M_*^{-1}$, and $\Lambda=0$. Circled are two points in
the $\phi'_{1}$-$\chi'_{1}$ parameter space with the same values
of $\mu_{2}^2$ and $\Sigma_{2}$, corresponding to two solutions
with the same Lagrangian (plotted in Fig.~\ref{bsoln1}).}
\label{contour1}
\end{figure}
\subsection{Example 2: Higher-Order Potential}
In this second example we take a slightly more complicated kink potential
\begin{equation}
V(\phi) = -\frac{1}{2}m_\phi^2\phi^2 - \frac{1}{4}\lambda\phi^4
+ \frac{1}{6}\xi\phi^6\ .
\end{equation}
The other potentials and boundary conditions are the same as in the
previous example, the only difference being the dynamical evolution
of the system due to the new potential $V(\phi)$. We choose this potential because, contrary to the
potential in our first example, in the limit of weak
gravity and flat spacetime, it leads to multiple solutions to the
same boundary value problem~\cite{Toharia:2007xe,Toharia:2007xf} due to the nonlinear nature
of the equations.
In our more general setting, including gravity and a stabilization field, we find numerically
that there exists more than one solution for the same Lagrangian in a large portion of the
parameters space. In Fig.~\ref{bsoln2} we show two such solutions.
\begin{figure}[t!]
\includegraphics[width=16cm,height=12cm]{bsoln2.eps}
\vspace{0.2cm}
\caption{Profiles of the scalar backgrounds $\phi(y)$ and $\chi(y)$
as well as the warp factor $\sigma(y)$, showing the two possible solutions
(panels A and B) to the same boundary value problem defined by
the physical parameters
$m_{\chi}^2=-0.5M_*^2$,
$\mu_{1}^2=-0.25M_*^2$, $\mu_2^2=-5M_*^2$, $\Sigma_{1}=-2M_*^4$, $\Sigma_2=0.56M_*^4$, $m_{\phi}^2=0.5M_*^2$,
$\lambda=2M_*^{-1}$, $\xi=6M_*^{-4}$, and $\Lambda=0$.}
\label{bsoln2}
\end{figure}
Note that these solutions would be extremely difficult to discover by randomly
guessing initial-boundary conditions. To be more methodical we follow
the same procedure as before to find level-curves of the
final-boundary conditions, shown in Fig.~\ref{contour2}. Again,
solutions to a particular action will be given by
the intersection of the appropriate contours for the brane mass squared
$\mu_{2}^2$ and brane tension $\Sigma_{2}$. As can be seen, there are
regions in which some contours intersect at more than one point,
showing that multiple solutions for the same action
are possible as expected. In particular, we again circle two such points,
corresponding to the solutions plotted in Figs.~\ref{bsoln2}. As an
interesting remark, note that both of these particular solutions happen
to lie near the region of parameter space where the 4D cosmological
constant vanishes.
\begin{figure}[t!]
\includegraphics[width=16cm,height=12cm]{contour2.eps}
\vspace{0.2cm}
\caption{Level-curves of $\mu_{2}^2$ and $\Sigma_{2}$ in the
$\phi'$-$\chi'$ parameter space for example 2 with
$m_{\chi}^2=-0.5M_*^2$, $\mu_{1}^2=-0.25M_*^2$,
$\Sigma_{1}=-2M_*^4$, $m_{\phi}^2=0.5M_*^2$,
$\lambda=2M_*^{-1}$, $\xi=6M_*^{-4}$ and $\Lambda=0$. Circled are
two points in the $\phi'_{1}$-$\chi'_{1}$ parameter space with the
same values of $\mu_{2}^2$ and $\Sigma_{2}$, corresponding to two
solutions with the same Lagrangian. These solutions are plotted in
Fig.~\ref{bsoln2}.}
\label{contour2}
\end{figure}
\section{Stability of Solutions}
\label{warpedxd_stability}
Having shown how different nontrivial static field configurations
exist in warped extra dimensions, the next question to ask is whether
these solutions are stable. As we reviewed in section~\ref{flatxd}, in the case
of flat extra dimensions there exist~\cite{Toharia:2007xf}
techniques for determining the stability of such solutions. Indeed, for certain potentials, in that case perturbative stability can be determined
analytically. Unfortunately, in the case of warped extra dimensions,
the question of stability is complicated by the presence of multiple
scalar fields and their coupled dynamics. Here we begin to study the perturbative stability of these kinked configurations. We
derive the linearized equations and reformulate the problem in terms
of a matrix Sturm-Liouville problem. However, the full analysis
requires matrix Sturm-Liouville methods which we omit and leave
for future work.
We begin by expanding the metric to first-order. Instead of the
coordinates in~(\ref{warped_metric}), in this section it will be more
convenient to choose coordinates so that the metric takes the form
\begin{equation}
ds^2 = a^2(y)\left(\gamma_{\mu\nu}(x)dx^\mu dx^\nu - dy^2\right) \ .
\label{conformal_warped_metric}
\end{equation}
Working in the generalized longitudinal gauge (see appendix~\ref{ScalarPerts} for details), we
introduce scalar perturbations $\Phi$ and $\Psi$ and write the perturbed metric as
\begin{equation}
ds^2 = a^2(y)\left[(1+2\Phi(x,y))\gamma_{\mu\nu}(x) dx^\mu dx^\nu
- (1+2\Psi(x,y))dy^2\right] \ .
\label{pert_metric}
\end{equation}
Next we expand the ${\cal N}$ scalar fields to first-order in small perturbations $\pchi_{a}(x,y)$
\begin{equation}
\chi_{a}(x,y)
= \bchi{}_{a}(y) + \pchi_{a}(x,y)\ ,
\end{equation}
and compute the linearized Einstein equations, yielding ${\cal N}+1$
dynamical equations for ${\cal N}+1$ scalar fields (${\cal N}$ fundamental scalars
and one graviscalar, or radion). Since only ${\cal N}$ of these equations are
independent, the Einstein constraint equations are used to
eliminate one of the scalar fields in terms of the others (see
Appendix~\ref{ScalarPerts} for details). The resulting ${\cal N}$
independent equations are
\begin{eqnarray}
{}^{(4)}\Box\Psi - \Psi''
- \left(9{\cal H} - 2a^2\frac{1}{\bchi_{{\cal N}}'}\left.\frac{\partial W}{\partial\chi_{{\cal N}}}\right|_{\bchi}\right)\Psi'
&-& \left(12{\cal H}^2 + 4{\cal H}' - \frac{1}{2}{}^{(4)}{\cal R} - 4a^2{\cal H}\frac{1}{\bchi_{{\cal N}}'}\left.\frac{\partial W}{\partial\chi_{{\cal N}}}\right|_{\bchi}\right)\Psi \nonumber \\
&=& -\frac{4a^2}{3M_*^3}\displaystyle\sum_{a=1}^{{\cal N}-1}
\left(\left.\frac{\partial W}{\partial\chi_{a}}\right|_{\bchi}
- \frac{\bchi_{a}'}{\bchi_{{\cal N}}'}\left.\frac{\partial W}{\partial\chi_{{\cal N}}}\right|_{\bchi}\right)\pchi_{a}
\label{pert_einstein}
\end{eqnarray}
\begin{equation}
{}^{(4)}\Box\pchi_{a} - \pchi_{a}'' - 3{\cal H}\pchi_{a}'
- a^2\left.\frac{\partial^2 W}{\partial\chi_{a}^2}\right|_{\bchi}\pchi_{a}
= - 3\bchi_{a}'\Psi'
- 2a^2\left.\frac{\partial W}{\partial\chi_{a}}\right|_{\bchi}\Psi\ ,
\label{pert_field}
\end{equation}
where ${\cal H}\equiv\frac{a'}{a}$, ${}^{(4)}\Box\equiv\gamma^{\mu\nu}\partial_{\mu}\partial_{\nu}$, and in equation (\ref{pert_field}), as
throughout, we have assumed that there are no direct couplings between
the 5D scalar fields in the scalar potential (in
Appendix~\ref{ScalarPerts} we derive the general form of these
equations when couplings between the fields are included). These
dynamical equations can be written more compactly as
\begin{eqnarray}
\Box\Psi + {\cal D}^y_{1}\Psi &=& {\cal D}^y_{2}\pchi
\label{compact1}
\\
\Box\pchi + {\cal D}^y_{3}\pchi &=& {\cal D}^y_{4}\Psi\ ,
\label{compact2}
\end{eqnarray}
where $\pchi$ has suppressed discrete indices which run over the
${\cal N}-1$ fundamental scalar fields, $\Psi$ is the graviscalar, the
${\cal D}^y_{i}$ are $y$-dependent differential operators (i.e., linear
differential operators having $y$-dependent coefficients and acting
only on functions of $y$) also with suppressed discrete indices, and
$\Box\equiv\gamma^{\mu\nu}\partial_{\mu}\partial_{\nu}$ is the 4D wave operator.
The boundary conditions are determined by integrating the equations of
motion across each brane. Integrating equations (\ref{pert_einstein})
and (\ref{pert_field}), these are found to be
\begin{eqnarray}
[\Psi']_{y_{i}}
- \left.2a^2\frac{1}{\bchi_{{\cal N}}'}
\left.\frac{\partial W}{\partial\chi_{{\cal N}}}\right|_{\bchi}\Psi'\right|_{y_{i}}
&-& 4\left.a^2{\cal H}\frac{1}{\bchi_{{\cal N}}'}
\left.\frac{\partial W}{\partial\chi_{{\cal N}}}\right|_{\bchi}\Psi\right|_{y_{i}}
\nonumber \\
&=& \left.\frac{4a^2}{3M_*^3}\displaystyle\sum_{a=1}^{{\cal N}-1}
\left(\left.\frac{\partial W}{\partial\chi_{a}}\right|_{\bchi}
- \frac{\bchi_{a}'}{\bchi_{{\cal N}}'}\left.\frac{\partial W}{\partial\chi_{{\cal N}}}\right|_{\bchi}\right)\pchi_{a}\right|_{y_{i}}
\label{pert_einstein_jxn}
\end{eqnarray}
\begin{equation}
[\pchi_{a}']_{y_{i}}
+ \left.a^2\left.\frac{\partial^2 W}{\partial\chi_{a}^2}\right|_{\bchi}
\pchi_{a}\right|_{y_{i}}
= 2\left.a^2
\left.\frac{\partial W}{\partial\chi_{a}}\right|_{\bchi}\Psi\right|_{y_{i}}\ .
\label{pert_field_jxn}
\end{equation}
These boundary conditions can be put in the form
\begin{eqnarray}
\Psi'(x,y_{i}) &=& A_{1}(y_{i}) \Psi(x,y_{i}) + A_{2}(y_{i}) \pchi(x,y_{i})
\label{Psibc}
\\
\pchi'(x,y_{i}) &=& B_{1}(y_{i}) \Psi(x,y_{i}) + B_{2}(y_{i}) \pchi(x,y_{i})\ ,
\label{pchibc}
\end{eqnarray}
where $A_{1,2}$ and $B_{1,2}$ are functions of $y_{i}$, defined via~(\ref{pert_field_jxn}), and
we have used~(\ref{Psibc}) in~(\ref{pert_einstein_jxn}) to obtain~(\ref{pchibc}).
The plan now is to perform a separation of variables in order to obtain a Sturm-Liouville
eigenvalue problem, and then to analyye this eigenvalue problem to determine
stability of the system. Because the 5D equations of motion of the
scalar perturbations are coupled, the correct separation of variables
ansatz is a coupled one
\begin{eqnarray}
\Psi^{(n)}(x,y) &=& \Psi_{y}^{(n)}(y)\ \ux^{(n)}(x)
\label{twist1}
\\
\pchi^{(n)}(x,y) &=& \pchi_{y}^{(n)}(y)\ \ux^{(n)}(x)\ ,
\label{twist2}
\end{eqnarray}
where $\ux^{(n)}(x)$ is the $n^{th}$ 4D Kaluza-Klein physical mode and
$\Psi_{y}^{(n)}(y)$ and $\pchi_{y}^{(n)}(y)$ are the wave functions. Plugging this ansatz into equations
(\ref{compact1}) and (\ref{compact2}) leads to the following coupled
equations
\begin{eqnarray}
\Psi_{y}(y)\Box \ux(x) + \ux(x){\cal D}^y_{1}\Psi_{y}(y)
&=& \ux(x){\cal D}^y_{2}\pchi_{y}(y)
\\
\pchi_{y}(y)\Box \ux(x) + \ux(x){\cal D}^y_{3}\pchi_{y}(y)
&=& \ux(x){\cal D}^y_{4}\Psi_{y}(y)\ .
\end{eqnarray}
The separation of variables thus yields a 4D wave equation for $\ux(x)$
\begin{equation}
\Box\ux(x) + m_{\ux}^2\ux(x) =0
\end{equation}
and a system of two coupled differential equations
\begin{eqnarray}
{\cal D}^y_{1} \Psi_{y}(y) - m_{\ux}^2 \Psi_{y}(y) &=& {\cal D}^y_{2}\pchi_{y}(y)\label{coupledeq1} \\
{\cal D}^y_{3} \pchi_{y}(y) - m_{\ux}^2 \pchi_{y}(y) &=& {\cal D}^y_{4}\Psi_{y}(y)
\label{coupledeq2}
\end{eqnarray}
with boundary conditions for the profiles
\begin{eqnarray}
\Psi_{y}'(y_{i}) &=& A_{1}(y_{i})\Psi_{y}(y_{i}) + A_{2}(y_{i})\pchi_{y}(y_{i})\\
\pchi_{y}'(y_{i}) &=& B_{1}(y_{i})\Psi_{y}(y_{i}) + B_{2}(y_{i})\pchi_{y}(y_{i})\ .
\end{eqnarray}
The system of equations (\ref{coupledeq1}) and (\ref{coupledeq2})
constitute an eigenvalue problem. The stability of the static
background around which we have added scalar perturbations
therefore depends on the existence, or absence, of a negative eigenvalue
$m_{\ux}^2$ associated with a solution to eqs.~(\ref{coupledeq1}) and
(\ref{coupledeq2}).
This situation is somewhat unusual, since generally the Kaluza-Klein
eigenvalue problem arising from dimensional reduction consists of a single
second order differential equation, which can be put in
standard Sturm-Liouville form. Analyzing that Sturm-Liouville
eigenvalue problem is straightforward, since in particular it is known that the
eigenvalues are bounded from below, and that the eigenfunction
corresponding to the smallest eigenvalue has no zeros
within the interval. Therefore, the question of stability in practical
terms becomes the search for a solution to the Kaluza-Klein equation
such that it contains no nodes. Its associated eigenvalue will be the
lightest possible eigenvalue and, if positive, the system will have no classical instabilities.
In the present case, however, the Kaluza-Klein problem is a system of coupled
differential equations. Consequently, matrix Sturm-Liouville
techniques are required. In order to analyze stability further, one
must extend the theory of oscillations and the concept of nodes of solutions to a higher
dimensional problem. Such an analysis, although rather involved, is underway, and will be presented in a
future work.
\section{Discussion and Outlook}
Braneworld theories generally lead to scalar degrees of freedom that
propagate in the extra-dimensional bulk. Understanding the vacuum
structure of these models in the presence of bulk scalar
fields is therefore a prerequisite to fully appreciating their
phenomenological possibilities. Furthermore, bulk scalars may provide a useful way to localize fermions and build braneworld models purely with field theory (e.g., fat branes and soft walls).
In this work, we have studied the vacuum structure of braneworld
models with one warped extra dimension and multiple bulk scalar
fields. In particular we have focused on static configurations along the
extra space coordinate where one of the fields--with Dirichlet
boundary conditions--acquires a nontrivial kink-like profile. To find
these solutions one needs to solve both the Einstein and the scalar
field equations. In the limit of a flat 5D metric
and weak gravity such solutions are known to exist, and the problem of
finding all possible static configurations as well as determining
their perturbative stability has been addressed and
solved~\cite{Toharia:2007xe,Toharia:2007xf}. Here we have built upon
this previous work to determine how warping along the extra dimension
effects the existence and stability of these kink-like solutions.
When considering a fixed warped background, it was sufficient to look
for nontrivial solutions for a single scalar field. In this case,
neglecting any backreaction of the scalar field on the gravitational
dynamics, we found that such kink-like solutions do indeed exist. As
in the case of a flat extra dimension, we were able to prove that any
kink-like solution with nodes in the bulk is unstable. Thus we have focused on
nodeless kink solutions and the trivial solution. However, in contrast to the flat case, in the presence
of warping we were only able to find a sufficient condition for
determing the stability of these solutions. We were therfore unable to
analytically determine stability for nontrivial solutions in a warped
background, even when that background is fixed (e.g., in
the Randall-Sundrum model with no backreaction). Instead we were forced to determine
stability numerically.
Including the dynamics of the gravitational sector forces the inclusion of additional scalar fields whose purpose is to
stabilize the size of the extra dimension. In that case we were again
able to find nontrivial kink-like configurations, except now for a
coupled multiple-field system. We have described a general graphical
technique to find all possible static configurations of the background
equations with one kink scalar field and an arbitrary number of
additional ``stabilization'' fields. The technique amounts to
generating solution surfaces by varying the shooting parameters needed
to solve the coupled system of equations. This technique also allows
us to look for multiple solutions with the same action. We have
demonstrated how to implement this technique in two simple examples,
where we considered one kink field and one stabilization field in the
presence of gravity. As in the flat case, when the potential for the
kink field is a higher-order polynomial (leading to higher-order
nonlinearity in the field equations), we found that multiple solutions
may exist for the same action. Interestingly, however, we also found
multiple solutions for the same action when the kink potential was a
fourth-order polynomial, which differs from the result obtained in a flat background.
We have addressed the issue of stability only partially. We have derived
the full 5D perturbative equations, including gravitational
perturbations, for multiple scalar fields in the presence of a warped
extra dimension. The system of equations constitute a matrix
eigenvalue problem, which must be analyzed using an extension of the
usual theorems coming from oscillation theory or Sturm-Liouville
eigenvalue problems. Such techniques exist in the mathematical
literature but due to the complexity of the task, we have left the
numerical analysis of the general case for a later work.
\acknowledgments
The work of M. Trodden and EJW is supported in part by National
Science Foundation grant PHY-0930521, by Department
of Energy grant DE-FG05-95ER40893-A020 and by NASA ATP grant NNX08AH27G.
M. Trodden is also supported by the Fay R. and Eugene L. Langberg Chair.
|
\section{Introduction}
\begin{figure*}
\centering
\includegraphics[width=\textwidth]{two_colour_ha_2_3.eps}
\caption{Two colour image of NGC 1275. The green is a broad band B filter (F435W) and the red a broad band R filter (F625W) with the subtracted scaled green filter (F550M) image removing the contribution from the galactic continuum. The two regions of extended star formation discussed in this paper, the Blue Loop and the Southern
filament are labelled.}
\end{figure*}
Brightest Cluster Galaxies (BCGs), the most massive galaxies known, offer the
opportunity to observe the heating and cooling processes in the Intra-Cluster
Medium (ICM) and provide the ideal environment to test theories of massive
galaxy formation.
Optical line-emitting nebulae surround about a third of all BCGs
and are found specifically in those where the cluster exhibits a strongly peaked
X-ray surface brightness profile and a cool, high density core, so called, `cool
core' clusters (see \citealt{crawford2004} for a discussion of the optical properties
of these clusters).
The mass of cool gas implied by the high X-ray luminosity (i.e. the energy loss rate of the gas) in the centre of cool core clusters is ten or more times the mass inferred from soft X-rays ($<1\,\mathrm{keV}$), observed star formation rates and the mass of cold gas present \citep{peterson2006}.
In order for the hot gas to be radiating but not cooling in the predicted quantities
there must be a source of heat regulating the production of cool gas.
Energy arguments favour feedback from a SuperMassive Black Hole (SMBH), situated in the central galaxy, as the dominant heating process in these clusters (for a review see \citealt{peterson2006, mcnamara2007}). Understanding the role the black hole plays with respect to the galaxy's environment such as
its connection with star formation, is a key question in galaxy evolution.
NGC 1275 is the BCG in the nearby ($z=0.0176$) Perseus Cluster, the X-ray brightest, cool core cluster. The NGC 1275 system is very complex;
\cite{minkowski1955} discovered it
consists of two galaxies, a high velocity system (HVS, $8200\,\mathrm{km}\,\mathrm{s}^{-1}$) and a low velocity system (LVS, $5200\,\mathrm{km}\,\mathrm{s}^{-1}$). 21 cm observations, line absorption in HI and Ly$\alpha$ and continuum absorption in X-rays \citep{ekers1976, briggs1982, fabian2000, gillmon2004} show the HVS to lie in front of the LVS. The HVS can be clearly seen in Fig. 1, north-west of the centre.
Early observations showed a spatial correspondence between the HVS and LVS \citep{hu1983, unger1990} and evidence for gas at intermediate velocities \citep{ferruit1997}, which led to the suggestion that the galaxies were in the process of merging. However, these observations can also be explained by influences of the ICM \citep{fabian1977, boroson1990, caulet1992} or by a past interaction, of one of the systems, with a third gas-rich galaxy \citep{holtzman1992, conselice2001}. By examining the X-ray absorption, \cite{gillmon2004} were able to put a lower limit on the distance between the HVS and the nucleus of the LVS, of $57\,\mathrm{kpc}$. This analysis was repeated with a deeper study by \cite{sanders2007} giving an improved lower limit on the distance between the two galaxies as $110\,\mathrm{kpc}$. This large separation and lack of obvious shocked gas suggests that any interaction between these two systems lies in their future not in their past.
NGC 1275 exhibits interesting optical features in both its relatively blue
central colours \citep{holtzman1992} indicating the presence of massive, short-lived stars and in its vast
extended emission-line nebulae \citep{minkowski1957, lynds1970} stretching out predominantly radially from the nucleus of the LVS. Large quantities of molecular hydrogen and CO gas has been found in the nuclear region \citep{lazareff1989, inoue1996, donahue2000} and recently detected in the outer H$\alpha$ filaments \citep{hatch2005, salome2008a, lim2008, ho2009}. Soft X-ray emission has also been found to be associated with some of the optical emission nebulae \citep{fabian2003}, however is less luminous than the optical and UV emission.
\cite{kent1979} found the low resolution spectra of the H$\alpha$ filaments to be similar to those of H II regions and well explained by collisional ionisation or photoionisation from the Seyfert nucleus. Their observations supported a hypothesis that the filaments were formed by an accretion flow onto the LVS. It has since been shown, with higher resolution spectra, that the galaxy nucleus is not the ionisation source \citep{johnstone1988}.
\cite{heckman1989} observed both the filaments in NGC 1275 and broadened their study to include filamentary systems in other BCGs. They found evidence confirming the association of optical emission-line nebulae and cooling flows and discuss mechanisms for heating and ionising these nebulae.
The central star cluster population in NGC 1275 has been well studied;
\cite{shields1990} detected H II regions and excess blue continuum in the central filaments of NGC 1275 which they interpreted as implying the presence of massive young star clusters. Soon after, \cite{holtzman1992} discovered a population of
compact massive blue star clusters in the core of NGC 1275 with the Hubble
Space Telescope (HST) Wide
Field Planetary Camera 1 (WF/PC1). The observed sizes and luminosities led the authors to conclude
that these were
most likely proto-globular clusters. The conclusion was supported by \cite{richer1993} using observations from the Canada-France-Hawaii Telescope (CFHT) and by \cite{carlson1998}
using HST WFPC2 observations. Assuming a Salpeter
Initial Mass Function (IMF), previous ages of the central clusters are suggested to be between $10^{8}-10^{9}$ years
resulting in masses between $2\times10^{7}-10^{8}\,\mathrm{M}_{\odot}$.
A further study of the colour distribution of the NGC 1275 system was carried out by \cite{mcnamara1996}.
The $U-I$ colours indicate a centrally concentrated young population and a more diffuse older background population.
The young population colours are consistent with ages up to $1\,\mathrm{Gyr}$. Measuring the physical parameters
of star clusters in cD galaxies can allow
us to constrain both their rate of formation and the underlying physical processes
behind their development.
The complicated nature of the NGC 1275 system makes it difficult to determine whether the star cluster populations in the center lie
in the LVS or HVS. \cite{keel2001} suggested that the correlated spatial distribution of the blue clusters and the HVS dust lanes are
evidence of this cluster population being part of the foreground HVS. However, \cite{brodie1998} have shown that at least some of the
blue clusters are at the same redshift ($\sim5200\,\mathrm{kms}^{-1}$) as the LVS. The colours of the blue population have been
determined to be similar, regardless of whether the individual clusters have been confirmed to lie in the LVS or are perhaps, spatially,
more likely to lie in the HVS \citep{holtzman1992, norgaardnielsen1993, richer1993, carlson1998}. \cite{dixon1996} observed NGC 1275 with
the Hopkins Ultraviolet Telescope and found evidence for a star formation rate of 30$\mathrm{M}_{\odot}\,\mathrm{yr}^{-1}$ at the redshift of the HVS.
There are two major competing scenarios for the formation of Globular Clusters (GC) in
elliptical galaxies.
\cite{ashman1992} propose that elliptical galaxies form through
mergers which trigger the GC formation while \cite{forbes1997} proposes that the GCs form `in situ' via collapse into a single potential well (for a review see \citealt{brodie2006}).
The merger
scenario produces a bi-modal distribution of clusters and is supported by the observation
that elliptical galaxies tend to have a higher specific frequency (number of clusters per unit galaxy light) than spiral
galaxies. The formation processes of the clusters in the centre of NGC 1275
are discussed by \cite{carlson1998} and \cite{richer1993}. The former
prefer a model where the star formation is triggered by a merger while the latter suggest the star formation is due to cool gas collected at the centre of the cooling flow. The
main contention is whether the colour-magnitude diagram shows evidence
of a single age population or of a continuously forming population of star clusters.
\cite{ferruit1994} used TIGER, an Integral Field Spectrograph (IFS) mounted on the $3.6\,\mathrm{m}$ CFHT to detect emission line regions coincident with the central clusters, previously assumed only continuum sources. These regions exhibit spectral features very different to the H II regions discovered previously and have spectral and kinematical properties characteristic of the filaments. The line ratios suggest that photoionisation by the star clusters cannot account for the gas ionisation in the filaments. These findings, coupled with the uncertainties of our knowledge of the internal reddening in NGC 1275, led the authors to support the theory that the clusters were formed from a cooling flow.
Further IFS observations with GMOS on GEMINI were taken by \cite{trancho2006}. These observations confirmed the discovery of emission lines associated with some star clusters, although the majority exhibit very little gas emission.
The star formation is however not limited to the central region (see regions marked on Fig. 1). NGC 1275
also exhibits regions of very extended star formation noted by \cite{sandage1971}. The `blue knots' of \cite{sandage1971} form the eastern side of the Blue Loop, approximately $22\,\mathrm{kpc}$ to the south-east of the nucleus (Fig. 1). \cite{conselice2001} presented colours and magnitudes for clusters
near the north-west of the HVS and a few bright clusters on the eastern arm of the Blue Loop. On the Blue Loop they measure very
blue colours with ($B-R$)$_{0}$ $\approx$ $-0.3$ to 0.0.
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{cover_2_2_final.eps}
\caption{Image showing the star clusters (blue) and the H$\alpha$ emission (red) that make up the Blue Loop. Image from \protect \cite{fabian2008}.}
\end{figure}
These objects can be
dated from their colour-magnitude relationships and through this we can examine their
relationship to the H$\alpha$ filaments and to the wider cooling and heating
flows in the galaxy.
Observations of cooling flows and star formation in BCGs can give us an
insight into the evolutionary cycle that the gas undergoes in these objects.
\begin{figure*}
\centering
\includegraphics[width=0.7\textwidth]{NGC1275_mag_col_0609.eps}
\caption{Vega magnitude (F435W$_{0}$) \protect \footnotemark against colour (F435W-F625W)$_{0}$ for the three regions. All data points have
errors in single passbands less than 0.15 mag. Panel (a) (Top) shows the
Blue Loop region (RA $03^{h}19^{m}52^{s}.1$, Dec. $41^{\circ}$ 30'08.6'' J2000), panel (b) the Southern filament (RA $03^{h}19^{m}46^{s}.5$, Dec. $41^{\circ}$ 30'04.5'' J2000) and panel (c) the south-west section (RA $03^{h}19^{m}47^{s}.9$, Dec. $41^{\circ}$ 30'38.8'' J2000) of the nucleus. The blue and red lines in panels (a) and (b) indicate the mean colours of the blue and red populations respectively. The blue and red populations are determined by inspection of the histogram (see Fig. 4). The lines in panel (c) show the average colours determined by \protect \cite{carlson1998} for their blue and red populations, in the same region.}
\end{figure*}
In this paper we present observations, obtained with the HST Advanced Camera for Surveys (ACS) spanning an
area of approximately 5.5 square arcmin. The observations allow us to analyse the star formation spatially coincident with the H$\alpha$ emission line filaments in the outer regions of the system. The regions are too far from the centre of NGC 1275 to be seen on the WFPC2 images of \cite{carlson1998}. In Section 2 we discuss the observations and data
reduction procedure and in Section 3 present the results of the analysis and
highlight potential sources of error. We discuss the results in the context of
different formation scenarios of the extended star formation in Section 4 and conclude in Section 5.
\begin{figure*}
\centering
\includegraphics[width=0.8\textwidth]{NGC1275_bl_hist_0609.eps}
\caption{(Left) The extinction-corrected colour-magnitude diagram for the Blue Loop. (Right) The number of clusters, in this region, detected in colour bins of width 0.1 mag. Here the bimodal distribution of colours can be clearly seen. The blue population (lower) has a broader distribution than the red population (upper), it also appears to have a tail towards the red end. The cut off between the two populations lies at $(\mathrm{F435W}-\mathrm{F625W})_{0} \sim 0.4$. The mean colour of the young, blue population is $(\mathrm{F435W}-\mathrm{F625W})_{0} = -0.22$.}
\end{figure*}
\section{Observations and Data Reduction}
\footnotetext{The $_{0}$ notation used for magnitudes and colours in this paper refer to corrections for foreground Galactic extinction only.}
Observations were made using the HST ACS on 2006 August 5th in the F435W, F550M and
F625W filters \citep{fabian2008}. Total exposure times were 9834, 12132 and 12405 s respectively.
The F625W filter includes H$\alpha$, [N II] and [S II] emission from both the HVS and the LVS. This differs from the \cite{conselice2001} ground-based, narrow-band imaging that excluded the HVS.
\begin{table}
\begin{center}
\begin{tabular}{|l|l|l|l|}
\hline
Filter & \multicolumn{3}{c}{Aperture Correction} \\
\hline \hline
& A & B & C \\
F435W & 1.07 & 1.06 & 1.05 \\
F550M & 1.07 & 1.06 & 1.04 \\
F625W & 1.08 & 1.07 & 1.04 \\
\hline
\end{tabular}
\end{center}
\caption{Aperture corrections for the three ACS filters used. A corresponds to the Blue Loop,
B the Southern filament and C the south-west portion of the nucleus. This correction was determined
from the 10 brightest sources in each region.}
\end{table}
\begin{table}
\begin{center}
\begin{tabular}{|l|l|}
\hline
Filter & Extinction Coefficient \\
\hline \hline
F435W & $A_{\mathrm{F435W}}=0.67$ \\
F550M & $A_{\mathrm{F550M}}=0.50$ \\
F625W & $A_{\mathrm{F625W}}=0.43$ \\
\hline
\end{tabular}
\end{center}
\caption{Galactic extinction coefficients for each ACS pass-band used.}
\end{table}
The standard procedure was used to bias-subtract and flat-field the data frames
which were then drizzled together as described in the HST ACS data handbook.
Fig. 1 shows the ACS F435W filter image of NGC 1275 combined with a red image made by subtracting the scaled F550M image from the broad band F625W filter, removing the contribution from the galactic continuum. Here the relation between the star formation and the line emission nebulae can be seen clearly. The HVS is in the line of sight of the nucleus of the LVS and appears to have a backward S shaped configuration.
Fig. 2 shows more clearly the relationship between
the H$\alpha$ filament and the star clusters in the Blue Loop region. Here it can be seen
that on the western arm the clusters are spatially offset from the filament by approximately 10 arcsec
corresponding
to a distance of 3.5\hbox{$\rm\thinspace kpc$}\ at a redshift of $z=0.0176$ (we adopt $H_{0}=71\,\mathrm{km}\,\mathrm{s}^{-1}\,\mathrm{Mpc}^{-1}$). Clusters on the eastern arm coincide directly (in the line of sight) with the H$\alpha$ filament.
We detected star clusters in the south-west region of NGC 1275 using IRAF \hbox{\rm DAOFIND}\ and the method of
\cite{carlson1998}. We find that the clusters in the Blue Loop and Southern filament region are more
compact than the clusters in the centre. The point-spread-function fitting technique
for crowded fields employed by IRAF \hbox{\rm DAOPHOT}\ was not used as many of the clusters are
partially resolved, and we required that a single technique should be used for
the whole field.
As in \cite{carlson1998}, we
required the \hbox{\rm DAOFIND}\ parameters of roundness between $-1$ and 1 and sharpness between 0.2 and 1. In the central image, only
the south-west region was considered; it is both less obviously `dusty' and
allows a more direct comparison of our colour-magnitude results with previous results.
\begin{figure}
\centering
\includegraphics[width=0.5\textwidth]{NGC1275_mag_col_0609_page3.eps}
\caption{Colour-magnitude diagram showing a comparison of the results for the Blue Loop and Southern Filament regions (in black) and for two control regions (in red) of size 15''$\times$15'' taken next to these regions. }
\end{figure}
{\sc IRAF} {\sc PHOT} was employed to perform aperture photometry using an aperture with a two-pixel radius in all regions
and all filters. The background in the Blue Loop ans Southern filament regions was estimated using the modal value in an annulus 10 pixels
wide with an
inner radius of 15 pixels. In all regions growth curves of the brightest, isolated star clusters in the field were
constructed to test the background estimation. The galaxy background near the outer regions of star
formation, far from the nucleus, does not vary much with position, however the field here is crowded. The large inner
radius of the background annulus insures we are sampling much more background than surrounding stars. In the central
region an inner radius of 12 pixels and annulus 2 pixels wide, as in \cite{carlson1998}, was found sufficient to estimate
the background. The small sky radius here is used to minimise the variation of the background due to the galaxy. No
subtraction of the smooth galaxy light in the inner region was done as the error introduced in the bright sources of
interest is small. The worst case scenario of a non-linear background variation over the 28 pixels, for the innermost
and therefore most steeply varying background introduces an error of less than 0.3 counts~s$^{-1}$. The variation of
background to peak for this innermost cluster is less than 2 per cent. In the inner region the bright galaxy nucleus
was masked out, as was a bright saturated star near the top of the eastern arm of the Blue Loop.
Average aperture corrections are determined using the 10 brightest sources in each of the three regions.
These are corrections from a two pixel radius aperture corresponding to
a 0.1 arc-second radius aperture flux to a flux determined with a 0.6 arc-second radius.
These corrections are listed in Table 1. The corrections are similar but slightly larger than
those found by \cite{carlson1998} in the core of the galaxy. The aperture correction for the three stellar regions was
determined separately and is found to be marginally larger in the outer regions. The point spread function changes with
position on the chip (errors of a few per cent), due to both optical aberrations and geometric distortions which may be
responsible for the small differences detected in aperture correction across the field.
Fig. 5 shows a colour-magnitude plot of the outer star formation regions accompanied by photometry of neighbouring control regions. We
do not find any objects as blue as our blue population of clusters in any of these fields, however the red population
appears ubiquitous throughout the field. Contamination of the young star cluster population from foreground and background
sources is therefore highly improbable due to both the lack of young blue sources in other regions and the spatial
distribution of the blue population in the Blue Loop and Southern filament regions. This is discussed further in section 3.1 and Fig. 7.
The ACS filters differ significantly from other filter systems and this needs
to be taken into account when determining reddening corrections from the galactic foreground.
Due to the differences in filter transmission curves between ground-based and the ACS filter systems
the extinction coefficients are calculated in the native photometric system and
all corrections applied before converting to another system.
For comparison with the stellar evolutionary synthesis models, we transform the ACS filter system to the Johnson-Cousins UBVRI system using the method below,
described in \cite{holtzman1995} and \cite{sirianni2005},
\begin{equation}
TMAG=SMAG+c0+c1 \times TCOL+c2 \times TCOL^{2}.
\end{equation}
The F435W filter can be transformed to a B band filter and F625W to a R band filter,
however uncertainties of a few per cent are introduced during this transformation. An additional systematic error is introduced when using the synthetic
transformation coefficients for stars with B$-$V$<0.5$ due to uncertainties in the shape of the total response curve in the F435W filter
(see \citealt{sirianni2005}, their Fig. 21). The overall uncertainty in these transformations is of the level 5-10 per cent.
For comparison Fig. 6
shows the B-band magnitude verses colour diagram for the
clusters after conversion to the Johnson-Cousins system. We used the synthetic transformation
coefficients given in Appendix D Table 22 in \cite{sirianni2005}
as these cover the larger colour range necessary for these results. We required the convergence
criterion to be $1\times10^{-5}$.
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{NGC1275_mag_col_JC.eps}
\caption{Johnson-Cousins $(B-R)_{0}$ colour against B-band Vega magnitude diagram for the Blue Loop region.}
\end{figure}
The extinction coefficients are determined by a constant times the reddening,
\begin{equation}
\frac{A_{\lambda}}{E(\lambda-V)}=R_{\lambda}.
\end{equation}
Here we have assumed
the constant, the ratio of total to selective extinction, is $R_{B}=3.1$ \citep{pei1992}.
This is the standard value derived from the $(B-V)_{0}$ colours in the Johnson-Cousins UBVRI band-passes,
using a ground based system.
The extinction ratios in the ACS filters were derived by \cite{sirianni2005} using the
\cite{cardelli1989} Galactic extinction law. The throughput in the medium and broadband
filters depends on the colour of the object. \cite{sirianni2005} selected three template
stars (05 V, G2 V and M0 V) from the Bruzual-Persson-Gunn-Stryker (BPGS) atlas of which we have adopted
the 05 V results to compute the extinction coefficients in the three filters (Table 2). The results are
normalised to the value $R_{B}=3.1$, however, there are systematic differences in extinction
between the ACS system and a ground based system as a function of $E(B-V)$ due to differences
in effective wavelength. The extinction in our ACS wide-band filters is systematically higher than
in the U B and V bands
as the effective wavelength is shorter. In the F814W
band the effective wavelength is longer and so the extinction here is systematically lower than in the I band \citep{sirianni2005}.
The value for reddening in the
direction of \hbox{\rm NGC 1275}\ we use is $E(B-V)=0.1627{\pm}0.0014$ determined using the Galactic reddening maps of \cite{schlegel1998}.
There will also be reddening from the dust in NGC 1275 itself. This will cause a greater
effect towards the centre of the galaxy, especially near the HVS, resulting
in the measurement of redder colours. The internal extinction is difficult to constrain, however
the two main areas of interest in this work, the Blue Loop and the Southern filament, are far from the dust lanes \citep{keel2001},
so we have not attempted to add a correction for this effect.
\section{Analysis and Results}
The H$\alpha$ filaments belonging to both the LVS and HVS have been investigated by \cite{conselice2001} (their Fig. 2.), and \cite{caulet1992} (their Fig. 2.) respectively. Their images show the extended filamentary system is predominantly associated with the large central galaxy not the smaller high velocity galaxy in the line of sight. \cite{hatch2006} identifies a 1-arcmin-long chain of blue star clusters in the north-west of NGC 1275 (RA $03^{h}19^{m}46^{s}.7$, Dec. $41^{\circ}$ 31'45.6'' J2000), an extension of those mentioned by \cite{conselice2001}. Only one blue knot within this chain appears to be spatially connected with the H$\alpha$ filaments. \cite{hatch2006} find that this blue knot has a line-of-sight velocity of 5538$\,$kms$^{-1}$ placing it within the LVS.
The following analysis assumes the star clusters along the two extended regions of star formation studied here also belong to the LVS (see also later discussion at end of section 4 and Fig. 14). Future spectroscopic data of these star clusters would allow us to place the clusters firmly within the HVS or LVS.
\begin{figure}
\centering
\includegraphics[width=0.5\textwidth]{blueloop_final_0609.eps}
\includegraphics[width=0.5\textwidth]{antenna_final_0609.eps}
\caption{Images showing spatial arrangement of blue and red clusters in both the Blue Loop (upper) and the Southern filament (lower). $x$ and $y$ axis here are both in pixels. In the Blue Loop region blue clusters are defined as those
with de-reddened $(\mathrm{F435W}-\mathrm{F625W})_{0} < 0.4$ and red clusters with de-reddened $(\mathrm{F435W}-\mathrm{F625W})_{0} > 0.4$. In the Southern filament blue clusters have $(\mathrm{F435W}-\mathrm{F625W})_{0} < 0.5$ and red clusters with $(\mathrm{F435W}-\mathrm{F625W})_{0} > 0.5$. These populations are determined by inspection of the colour histogram (see Fig. 4) in each region.}
\end{figure}
\subsection{Spatial Distribution of Clusters}
The clusters in the south-west region contain a range of colours evenly distributed about the core of the galaxy, implying the old and young population are inter-dispersed throughout the core of the LVS. The south-west region is spatially the most distinct region from the HVS in the core of NGC 1275, however confusion with this system can not be ruled out.
In the Blue Loop region all `bluer' clusters (defined as those with de-reddened $(\mathrm{F435W}-\mathrm{F625W})_{0} < 0.4$) are found to lie along the loop, which coincides with the H$\alpha$ emission (see Fig. 7). The `redder' clusters (de-reddened $(\mathrm{F435W}-\mathrm{F625W})_{0} > 0.4$) are evenly distributed and have the same colours as clusters found in a control region of the galaxy at a similar radius. The control region used has a central RA $03^{h}19^{m}54^{s}.1$ and Dec. $41^{\circ}$ 30'37.7'' (J2000) and was 10 by 10 arc-seconds across. This is also found to be the case in the Southern filament region, here blue cluster have $(\mathrm{F435W}-\mathrm{F625W})_{0} < 0.5$ and red clusters have $(\mathrm{F435W}-\mathrm{F625W})_{0} > 0.5$. Projection effects make it difficult to be precise about the arrangement of the extended filamentary systems in NGC 1275, however the close nature of the filaments and the stellar regions, and the fact that we see this apparent association in more than one area, yet the filaments themselves are very sparse in the extended regions, suggests that the star formation is likely to be intimately linked to the optically-emitting filaments.
\subsection{Stellar Populations Models}
We use Simple Stellar Populations (SSP) models to estimate the ages of the young populations of clusters. It is likely that these clusters formed almost simultaneously and with similar metallicities allowing them to be described by these models.
The commonly used models of \cite{bc03} (hereafter BC03) and the GALEV evolutionary synthesis models \citep{Kotulla2009} have been used to determine an upper bound on the ages of the star clusters
in NGC 1275. The GALEV models have the advantage of including the ACS filter system while the BC03 models require a photometric transformation to the Johnson-Cousins UBVRI system. This transformation introduces errors of a few per cent \citep{sirianni2005}.
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{figure_6.eps}
\caption{($\mathrm{F435W}-\mathrm{F625W})_{0}$ colour verses age relationship for several metallicities, measured relative to solar, using the GALEV SSP models, assuming a Salpeter IMF with lower mass cut-off $m_{L}=0.1\,\mathrm{M}_{\odot}$. The upper mass cut-off depends on metallicity with $m_{U}=50\,\mathrm{M}_{\odot}$ for super-solar and $m_{U}=70\,\mathrm{M}_{\odot}$ otherwise. The hashed lines show the spread in colour of the young, blue population in the Blue Loop region.}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{figure_7.eps}
\caption{The $(\mathrm{F435W}-\mathrm{F625W})_{0}$ colour verses age relationship using Galev SSP models assuming solar metallicity with a Kroupa, Salpeter and Scalo IMF respectively.}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{BC_col_age.eps}
\caption{Johnson-Cousins $(B-R)_{0}$ colour versus age relationships for three different metallicities using BC03 models for a single age, single burst population and assuming a Salpeter IMF. Lower and upper mass cut-offs are assumed to be $m_{L}=0.1\,\mathrm{M}_{\odot}$ and $m_{U}=100\,\mathrm{M}_{\odot}$ respectively. The hashed lines show the spread in colour of the young, blue population in the Blue Loop region.}
\end{figure}
The GALEV SSP models are shown for a range of metallicities and Initial Mass Functions (IMF's) ($\Phi(m)^{\alpha}$) in Fig. 8 and 9. These are all computed using the Padova 1994 isochrone models. \cite{Kotulla2009} assume a lower mass cut-off of $m_{L}=0.1\,\mathrm{M}_{\odot}$ while they allow the upper mass cut-off to depend on metallicity. A cut-off of $m_{U}=50\,\mathrm{M}_{\odot}$ is assumed for super-solar metallicity and $m_{U}=70\,\mathrm{M}_{\odot}$ otherwise. The Salpeter, Scalo and Kroupa IMF's assumed are defined with $\alpha$ given in equations 3, 4 and 5 respectively.
\begin{equation}
\alpha=-2.35
\end{equation}
\begin{equation}
\alpha \left\{
\begin{array}{rl}
-1.25 & \text{if } m \leq 1\,\mathrm{M}_{\odot}\\
-2.35 & \text{if } 1\,\mathrm{M}_{\odot} < m \leq 2\,\mathrm{M}_{\odot}\\
-1.25 & \text{if } 0.5\,\mathrm{M}_{\odot} < m
\end{array} \right.
\end{equation}
\begin{equation}
\alpha \left\{
\begin{array}{rl}
-1.30 & \text{if } m \leq 0.5\,\mathrm{M}_{\odot}\\
-2.30 & \text{if } 0.5\,\mathrm{M}_{\odot} < m
\end{array} \right.
\end{equation}
\\
The BC03 SSP models (Fig. 10) are computed assuming a Salpeter IMF
and the Padova 1994 evolutionary tracks. The models assume an IMF with a low mass cut-off
of $m_{L}=0.1\,\mathrm{M}_{\odot}$ and a high mass cut-off of $m_{U}=100\,\mathrm{M}_{\odot}$. The models are normalised
to a total mass of $1\,\mathrm{M}_{\odot}$ in stars at age $t=0$.
X-ray observations have shown the metallicity of the ICM to be approximately 0.6 of solar abundances at the Blue Loop region \citep{sanders2004, sanders2007}. Given this, metallicity abundances as high as twice solar or higher (Z$=0.05$), in these stars, are very unlikely as are abundances as low as 2 per cent (Z=0.0004) solar (Fig. 8). We might expect differences in the metallicities of the old and young cluster populations, the old clusters are potentially metal-poor globulars while the young clusters are likely to have metallicities $\sim$ solar. The SSP models used agree well for ages larger than 10$^{7}$ years, however modelling very young stars is complex and the models remain uncertain. This, coupled with errors on the data make it impossible to constrain the metallicities of the regions.
The observed colour of the clusters are highly sensitive to age, metallicity, dust extinction and the coincidence of the clusters with H$\alpha$ filaments. Older, metal-rich, dusty populations will appear redder. The star formation in regions such as the Blue Loop is far from the nucleus ($\sim22\,\mathrm{kpc}$), and exhibits no obvious signs of dust, so internal reddening is unlikely to have a large effect. As discussed above the metallicities are likely to be $\sim$ solar, however, as seen by Figs. 8 and 10, changes in metallicity can have a significant effect on the colour of young stellar populations. The clusters in our sample are coincident with H$\alpha$ emission. This will have the effect of reddening the colours of these clusters and may make them appear older than they are if the emission is within the photometry aperture or make them appear bluer if the emission is within the background annulus. There are ~twice as many counts in the brightest filamentary regions on the Blue Loop than counts in the background and for fairly bright sources (25 mag) the background counts are only ~2 per cent of the counts from the cluster. To make a 10 per cent difference to the $(\mathrm{F345W}-\mathrm{F625W})_{0}$ colour the H$\alpha$ (+[NII]) emission would require an equivalent width of $>100\,\AA$. This is very large and unlikely to be an issue.
\begin{figure}
\centering
\includegraphics[width=0.5\textwidth]{blue_loop_blue_small.eps}
\caption{The eastern arm of the Blue Loop showing that the clusters appear elongated possibly indicative of tidal stripping by the galaxy core.}
\end{figure}
\subsubsection{Blue Loop}
The bimodal distribution in colour shown in Fig. 4 strongly argues for a two-age population of clusters in this region. The colours of the `bluer' region are very blue with an average, extinction corrected colour of $(\mathrm{F435W}-\mathrm{F625W})_{0} = -0.2$, much bluer than the proto-globular clusters found in the core of the system ($\sim 0.3$). This also argues for a low internal extinction in this region.
Assuming a metallicity of $Z=0.02$ ($=Z_{\odot}$) for the star clusters and a colour range between $-0.4$ and $0.2$,
lower and upper age estimates from the BC03 SSP models are $5\times10^{6}$ and $5\times10^{7}$ years. The GALEV models yield an upper limit of $0.75\times10^{8}$ years.
The ages and luminosities can also give us estimates of the total mass in the clusters and hence
put limits on the star formation rates. We get a lower limit of $9\times10^{8}\,\mathrm{M}_{\odot}$
for all the blue clusters summed together. This provides a limit on the star formation of about
20 $\mathrm{M}_{\odot}\,\mathrm{yr}^{-1}$ over the whole Blue Loop region. The average mass in the clusters is found to be $4\times10^{6}\,\mathrm{M}_{\odot}$ with the brightest clusters having masses $\sim10^{7}\,\mathrm{M}_{\odot}$.
In the central region, \cite{carlson1998} found mass estimates for their brightest clusters of about $10^{7}-10^{8}\,\mathrm{M}_{\odot}$.
These results are highly dependent on the IMF and the models assume
a single burst, single age population. In the Blue Loop, the relatively large amount of scatter in panel (a)
in Fig. 3 suggests, assuming these clusters all share the same metallicity, that it does not have a true single age population. This could be evidence of continuous star
formation which `turned on' approximately $10^{8}$ years ago, the likely dynamical lifetime of the H$\alpha$ filaments \citep{hatch2005}. However we note, for very young stellar populations ($<10^{7}$~years) the SSP models are uncertain and the breadth of the colour population may not necessarily translate into a large age difference. A stellar population with an age 10$^{6}-10^{7}$~years can span the whole range of observed colours using BC03 models while the GALEV SSP models cannot account for any of the bluest clusters observed.
There is also a large amount of diffuse blue light around the Blue Loop region and particularly
on the north-west side towards the galaxy. This is most likely emission from smaller clusters and
single stars, possibly tidally stripped, or dissolved from, surrounding larger clusters. If these stars are similar in age to the clusters in the loop we would expect the total mass in stars, and therefore also the star formation rate calculated, to be a lower limit.
\subsubsection{Southern Filament}
The Southern filament region shows slightly redder colours, $(\mathrm{F435W}-\mathrm{F625W})_{0}$ $\sim$ $-0.1$ to 0.2, than the Blue Loop region, $(\mathrm{F435W}-\mathrm{F625W})_{0}$ $\sim$ $-0.4$ to 0.1 with a lower and upper
limit on the ages of the star clusters of about $8\times10^{6}$ and $5\times10^{8}$ years from the BC03 models and an upper limit of $5\times10^{8}$ years from the GALEV models. Fig. 3 panel (b) shows there is a clear bimodal distribution in this region however the distribution of colours in the young blue population is again fairly broad.
\subsubsection{South-west Nuclear Region}
The age estimates from the models applied to the south-west portion of the nucleus give limits
of $10^{7}$ and $10^{9}$ years, in agreement with the values estimated by \cite{carlson1998} and with the spectroscopically determined ages of five of the brightest inner clusters by \cite{brodie1998}.
\begin{table}
\begin{center}
\begin{tabular}{|c|c|c|c|}
\hline
ID & F435W$_{0}$ & F625W$_{0}$ & $(\mathrm{F435W}-\mathrm{F625W})_{0}$ \\
\hline \hline
1 & 21.559$\pm$0.013 & 21.006$\pm$0.011 & $\,0.55\pm0.017$ \\
2 & 21.324$\pm$0.011 & 21.349$\pm$0.014 & $-0.03\pm0.018$ \\
3 & 21.801$\pm$0.016 & 21.697$\pm$0.019 & $\,0.10\pm0.025$ \\
4 & 21.083$\pm$0.008 & 21.088$\pm$0.011 & $-0.00\pm0.014$ \\
5 & 21.183$\pm$0.010 & 21.255$\pm$0.014 & $-0.07\pm0.017$ \\
6 & 21.220$\pm$0.010 & 21.218$\pm$0.013 & $\,0.00\pm0.016$ \\
7 & 21.992$\pm$0.025 & 21.958$\pm$0.032 & $\,0.03\pm0.041$ \\
8 & 21.393$\pm$0.014 & 21.439$\pm$0.018 & $-0.05\pm0.023$ \\
9 & 21.707$\pm$0.013 & 21.700$\pm$0.016 & $\,0.01\pm0.021$ \\
10 & 20.898$\pm$0.006 & 21.006$\pm$0.009 & $-0.11\pm0.011$ \\
\hline
\end{tabular}
\end{center}
\caption{Magnitudes and colours for apertures 1 to 10 shown in Fig. 12.}
\end{table}
\subsection{Tidal Forces}
Fig. 11 shows that there are `groups' of star clusters in the Blue Loop which have a structure
reminiscent of a tail. One explanation for this could be that when the stars have formed they are unable to escape the gravitational potential of the galaxy and fall
back in towards the the nucleus.
If the star clusters have been tidally stripped, we would expect a gradient in the ages of the blue population. The star clusters forming first would have been subject to tidal stripping for longer and so should have fallen further towards the galaxy.
\begin{figure}
\centering
\includegraphics[width=0.5\textwidth]{apatures_age_dif_NGC1275_3.eps}
\caption{Apertures with a 9 pixel radius to test whether there is any discernible age difference along the `groups' of star clusters. The results of the photometry are given in Table 3.}
\end{figure}
To detect whether there is a discernible age difference across the star cluster `groups' in the bright eastern arm of the Blue Loop we choose 5 prominent groups of clusters. These are shown in Fig. 12 and the results of the photometry are given in Table 3. Three of the regions, those with apertures 2/3, 8/7 and 10/9 are found to have the younger, bluer clusters towards the south and two, 1/6 and 4/5, the younger clusters to the north. The difference detected in both 8/7 and 4/5 is small compared with the error in colour. The detected difference in colour in region 1/6 is very large but on closer inspection of the apertures it is not clear that they cover the same group of clusters so this region should be discounted.
Region 2/3 and 10/9 show a difference in colour larger than the error (change of 0.4 and 0.12 between apertures respectively). In both regions the younger clusters appear to lie in the southern most aperture. This is expected if the star clusters have formed at different times and not in some rapid event.
The age of the clusters can be estimated by assuming each elongated group to be experiencing
tidal stripping as it falls into the nucleus. The groups have an initial angular momentum from
the group formation process. The ratio of the length to width ($\Delta s / \Delta r \approx 5$, see Fig. 11) of each of these groups
of clusters can then be measured and the time over which they have been
stripped equated with the age by making the assumption that as soon as the gas was able to condense
out of the filament and form stars they started to fall into the nucleus. The typical ratio of $\Delta s$ to $\Delta r$ is measured from the ACS images note though, that projection effects could significantly alter this ratio.
Equation 6 and 7 assume that the clusters within each of these groups started to form in one place so the spread of the group ($\Delta s$ and $\Delta r$) seen now is the result of the initial velocity of the gas and tidal forces from the attraction of the LVS.
The gravitational acceleration felt by the element is $g\approx v^{2}/r$ where $v=700\,\mathrm{km}\,\mathrm{s}^{-1}$ is the velocity
dispersion. The velocity dispersion is inferred from the temperature of the ICM \citep{fabian2006} measured in X-rays. The tidal acceleration is given by,
\begin{equation}
\Delta g \approx \frac{GM \Delta r}{r^{3}} \sim \frac{g\Delta r}{r}
\end{equation}
so after time $t$,
\begin{equation}
\frac{\Delta s}{\Delta r} = \frac{1}{2}\frac{v^{2}t^{2}}{r^{2}}
\end{equation}
The distance from the centre of the potential to the group is $r = 63''$ which is about $22\,\mathrm{kpc}$ at the
distance of NGC 1275. With the ratio $\Delta s/\Delta r \approx 5$ and a velocity dispersion of $v=700\,\mathrm{km}\,\mathrm{s}^{-1}$ we calculate an age of $9\times10^{7}$ years, approximately the same as the
upper bound on the estimate from the SSP models. The precise nature of the tidal distortion for either radial or tangential motion depends on $v(r)$ which in turn depends on the mass distribution.
\section{Discussion}
In this paper we present photometry results for star clusters close to the nucleus and at a radius of $\sim$ 1 arcmin in the dominant galaxy in the Perseus cluster, NGC 1275. We give limits on the ages and masses of the star clusters and find a marked difference in the ages of the stars in the outer regions compared to those closer to the nucleus of the galaxy. In this section we discuss the implications of these results for the formations scenarios of the inner and outer cluster regions in turn.
\subsection{Inner Stellar Regions}
Our results, corrected for galactic extinction but uncorrected for internal extinction, yield colours of $(\mathrm{F435W}-\mathrm{F625W})_{0}$ $\sim$ 0.2 to 0.6 for the bluer star cluster population in the centre of NGC 1275. From SSP models we derive lower and upper limits on the ages of the clusters as $10^{7}-10^{9}$ years respectively, and an upper limit on the mass of the brightest clusters of $10^{8}\,\mathrm{M}_{\odot}$.
\cite{holtzman1992} were the first to discuss the formation scenarios of the bimodal population of star clusters in the LVS in NGC 1275. They detected a central population of blue clusters and explored three possible formation processes: (1) An interaction with the HVS, (2) star formation triggered by the cluster cooling flow and (3) star formation from a previous interaction with a third gas-rich galaxy. They rule out (1) primarily due to the high velocity ($3000\,\mathrm{km}\,\mathrm{s}^{-1}$) at which the two systems are falling toward each other and the clusters being roughly symmetrically distributed around the nucleus, rather than having a preferential direction towards the HVS. This conclusion was also reached by \cite{unger1990}. The second scenario was ruled out by the authors on the basis of the observed uniformity of the colour of the young, blue population of clusters. The same grounds favour scenario (3). This latter interpretation is supported by HST WFPC2 observations made by \cite{carlson1998}. However, \cite{richer1993} found a broader range in colours with no obvious bi-modality in their colour-magnitude diagram, which could indicate a large range in cluster ages and favours a scenario whereby the star formation is a formed in a more continuous manner. This is supported spectrally by the work of \cite{ferruit1994} who found the gas coincident with some of the inner clusters to have spectral properties similar to the gas in the filaments.
Two issues have changed since the earlier work involving scenario (3). The first is that the apparent destruction of the HVS at over 100$\,\mathrm{kpc}$ out from NGC 1275 emphasises that the ICM can prevent gas-rich galaxies penetrating right to the centre. The second is the enormous abundance of molecular gas already at the centre (more than $5\times10^{10}\,\mathrm{M}_{\odot}$, \citealt{salome2006}) which can easily supply fuel for star formation if dragged out by bubbles.
\begin{figure}
\centering
\includegraphics[width=0.5\textwidth]{xray_contoursha2_2.eps}
\caption{X-ray image ($0.3 - 7.0\,\mathrm{keV}$) with overlaid H$\alpha$ contours.}
\end{figure}
In this work we find evidence for a bimodal distribution of colours in the star clusters located in the south-west portion of the nucleus. However we detect a larger scatter about the modal value than that found by \cite{carlson1998}. Difficulties in photometry caused by a highly varying background, uncertain internal extinction and confusion with the HVS, do not allow us to conclusively rule out a history of formation by an interaction/merger with another galaxy, or a history of extended star formation in the younger, bluer population of clusters in the centre of NGC 1275.
\subsection{Outer Stellar Regions}
In both the Blue Loop and the Southern filament we find much `bluer' colours than in the core, with the former being the `bluest'. Colours in the Blue Loop of $(\mathrm{F435W}-\mathrm{F625W})_{0}$ $\sim$ $-0.4$ to 0.1 and the Southern filament of $(\mathrm{F435W}-\mathrm{F625W})_{0}$ $\sim$ $-0.1$ to 0.2, yield upper limits on the age of the extended star formation of $5\times10^{7}$ and $10^{8}$ years respectively, a factor of 10 younger than the blue clusters in the central region.
\cite{conselice2001} found colours on the bright eastern wing of the Blue Loop to be in the range $(B-R)_{0}$ $\sim$ $-0.3$ to 0.0 consistent with our results. The very blue colours and the large distance from the nucleus ($\sim20\,\mathrm{kpc}$), found here suggests that the internal reddening in this region is small and argues strongly for a two-age population in these regions.
\subsection{Star Formation Scenarios}
We consider two formation scenarios for the outer stellar regions: (1) Formation related to a previous interaction assumed responsible for the inner stellar populations and (2) a formation scenario entirely independent of the young, inner stellar populations.
\subsubsection{(1) -- Star Formation Triggered by an Assumed Previous Interaction}
For star formation in the extended stellar regions to be triggered by a previous interaction, the young, blue populations should either be coeval or the interaction needs to have had a delayed effect causing the star formation in the inner regions to occur before the outer regions.
If the outer and inner regions of star formation are the same age there must be a large amount of internal extinction in the core of the galaxy. Assuming no internal extinction in the Blue Loop, the colour correction needed for this region to be of comparable age to the inner clusters is $E(B-V)$ $\sim$ 0.6. This means that the difference between the B and R band extinctions would need to be three times greater than the extinction calculated here to correct for Galactic reddening along the line of sight to NGC 1275.
There is significant internal reddening from the LVS and also from the HVS. \cite{shields1990} use $A_{V}=1.2$ which gives an $E(B-V) = 0.39$ using the typical ratio of the extinction to the reddening of 3.1. We minimise the reddening due to the HVS by observing only the south-west, central region that is least obviously connected with the HVS.
The spatial distribution and distance from the core, of the extended star formation and the linear nature of the H$\alpha$ filaments argue against both of these scenarios. A merger or interaction of some kind is unlikely to disturb the Blue Loop and Southern filament systems enough to cause large amounts of localised star formation in these regions without disturbing other filaments nearby. The range of colours, especially in the Blue Loop, is also quite large, this could mean that the population is not all the same age and therefore unlikely to have been formed in some rapid event.
\subsubsection{(2) -- Formation Scenarios Independent of an Assumed Previous Interaction}
The extended blue populations of star clusters are particularly
intriguing as they appear spatially distributed along or just offset
from some radial H$\alpha$ filaments but only in certain regions of
the galaxy; {\it most H$\alpha$ filaments are not accompanied by star
formation}. Another galaxy-galaxy
interaction is unlikely to be responsible for this young population for
both the reasons discussed above and due to the very young age of these
clusters, the oldest in the Blue Loop being at most $5\times$10$^{7}$~years.
With the HST images, \cite{fabian2008} have resolved
some filaments into a collection of threads, each of which has a
diameter of about 70~pc. They argue that magnetic fields close to the
pressure equipartition value of the hot gas, of about 100$\mu$G, are
required for the thin threads within the filaments to maintain their
integrity against tidal forces in the cluster core. Filaments with a
higher mean density would not be supportable. The general lack of
massive star formation within the filaments is then due to magnetic
fields which balance the self gravity of the cold gas (see
\citealt{mckee1993} for a review of the role of magnetic fields in
star formation and \citealt{ho2009} for a discussion on the inner
region of NGC\,1275).
\cite{fabian2008} find that a typical thread has a radius of 35~pc, a
length ($l$) of 6~kpc and a mass of about $10^6\mathrm{M}_{\odot}$ (obtained
by scaling the H$\alpha$ emission to a filament complex of total mass,
obtained from CO observations of $10^8\mathrm{M}_{\odot}$
\citealt{salome2008}). The mass of a whole 6~kpc long thread therefore
corresponds to the mass of one observed star cluster. The
perpendicular column density $N\sim 4\times 10^{20}\hbox{$\rm\thinspace cm^{-2}$}$ or
$\Sigma_{\rm \perp}\sim 7\times 10^{-4}\hbox{$\g\cm^{-2}\,$}$. The lengthwise column
density, $\Sigma_{\parallel},$ is $\ell/2r$ times larger. The critical
surface density $\Sigma_{\rm c}$ for gravitational instability
corresponding to the magnetic field inferred in the filaments is given
by $\Sigma_{\rm c}=B/2\pi\sqrt{G}=0.062 (B/100\mu G)\hbox{$\g\cm^{-2}\,$}$, which is
about two orders of magnitude larger than the inferred value for a
thread, which is therefore individually very stable against
gravitational collapse.
This leaves the question, why is there any star formation at all, and
why does it only occur in certain places? We can only speculate here
but suspect that stars form when and where threads knot and clump
together to make much larger structures. There are many examples in
the images of NGC\,1275 where the filaments appear knotty. If a region
is disturbed or stretched in a radial direction, so that the cold gas can accumulate, then
clumps which are gravitational unstable can arise and form the star
clusters which have been the subject of this work. Examples where gas
may be accumulating as the field lines are stretched out are the ends
of the horseshoe (for this feature see Fig.~13 and Fig.~3 of
\citealt{fabian2008}). There appears to be no star formation there at
present but H$\alpha$-bright knots are seen.
Clumps or parts of filaments which are gravitationally unstable have a
column density which is much too high to be ``pinned'' to the
surrounding hot gas by magnetic fields and presumably have their own
ballistic orbits. The process of star formation might then be
dynamical. Threads within the long-lived outer filaments are supported
by magnetic fields and are stable to gravitational collapse. Where the
gas aggregates then they separate from the hot gas and, if sufficiently
dense, collapse to form star clusters.
\begin{figure*}
\centering
\begin{center} \subfigure[]{\label{fig:edge-a}}\includegraphics[width=0.32\textwidth]{temp_map_bl_2.eps}
\subfigure[]{\label{fig:edge-b}}\includegraphics[width=0.32\textwidth]{B_contours_bl_2.eps}
\subfigure[]{\label{fig:edge-c}}\includegraphics[width=0.32\textwidth]{Ha_contours_bl_2.eps}
\end{center}
\caption{The Blue Loop region: (a) X-ray temperature map, the dark regions are temperatures of $\sim2\,\mathrm{keV}$ and the light colours indicate temperatures of $\sim4\,\mathrm{keV}$. (b) B-band contours. (c) Contours of H$\alpha$ emission. The images are all the same scale, a black X marks the centre of the Blue Loop region on each image.}
\label{fig:edge}
\end{figure*}
We can put a limit on the maximum mass of a cluster using the
arguments of \cite{mckee1993}, assuming a steady, poloidal, time
independent magnetic field. Roughly, the magnetic energy scales as
$\mathcal{M}\sim B^{2}R^{3}$ and the gravitational energy scales as
$W\sim M^{2}/R$ so for a given field there is a critical mass where
the magnetic field can no longer support the filament against
gravitational collapse. This mass, $M_{\rm B}$, for a magnetised cloud
of radius $R$ and magnetic field $B$, with a constant of 0.12
determined from numerical calculations allowing for deviations from
uniform spherical clouds \citep{tomisaka1988}, is
\begin{equation}
M_{\rm B}=0.12\times \frac{B R^2}{G^{1/2}}.
\end{equation}
Taking typical values for a thread, then $B R^2$ is
$\pi\times(35\,\mathrm{pc})^{2}\times100\,\mathrm{\mu G}$ resulting in
$M_{\rm B}$ at just under $10^{6}\,\mathrm{\mathrm{M}_{\odot}}$. This is just
less than the average mass of a young cluster in the Blue Loop region
of $4\times10^{6}\,\mathrm{M}_{\odot}$. In order for a thread-sized region to
achieve gravitational collapse its mean density needs to rise by about
x100 above what is inferred for a typical thread
(\citealt{fabian2008}). The cold material therefore needs to aggregate
considerably, as discussed above. Note that the mean density of a
resolved thread of $\sim 2\hbox{$\rm\thinspace cm^{-3}$}$ is much lower than the actual
density of the cold gas ($\sim 100-1000\hbox{$\rm\thinspace cm^{-3}$}$) if its thermal
pressure is comparable to the thermal pressure of the surrounding
X-ray emitting gas where $nT\sim 10^6\hbox{$\rm\thinspace cm^{-3}~K$}$. (in other words the
volume filling factor of cold gas in a thread is low.)
The brightest star clusters have masses on the order
$10^{7}\,\mathrm{M}_{\odot}$. This requires them to originate from regions
several times larger than a thread. Since such regions cannot be
suspended by the hot gas, then the observed values of the thickness of
long-lived threads no longer apply. As
mentioned above one explanation is that the threads knot together and
collapse along their lengths, allowing them to evolve from stable to
unstable structures over time.
We can only speculate on what has triggered to star formation of the
Blue Loop. It coincides with a hotter region in the X-ray temperature
map (Fig.~14), which is the best evidence that the stars are
associated with the LVS and not the HVS. Clear confirmation must await
spectroscopic measurements of the star clusters. The blue loop roughly
lies diametrically opposite the horseshoe and the outer ghost bubble
seen in X-ray maps, which could indicate that the inner jets once went
along a NW--SE axis rather than the current almost N--S one (for example,
the jets may precess \citealt{dunn2006}). There is a particularly
bright arc of X-ray emission in the SE direction between the Blue Loop and the nucleus, which means the X-ray
emitting gas is denser there. If this caused disruption of the jet
then the jet interaction may have been unusual and have led to more
cold gas being dragged out or the gas out there having been strongly
disturbed. We can find no particular reason why the Southern filament
shows star formation.
\section{Conclusions}
In this paper we present ACS observations of the extended star formation in NGC 1275, assumed to be associated with the low velocity system and tracing the spatial distribution of the H$\alpha$ filamentary system.
We find a bimodal distribution with average colours of $(\mathrm{F435W}-\mathrm{F625W})_{0}$ $\sim$ $-0.4$ to 0.1 and $(\mathrm{F435W}-\mathrm{F625W})_{0}$ $\sim$ $-0.1$ to 0.2 for the blue population in the Blue Loop and the Southern filament respectively. In the central region we find much redder colours, consistent with previous results. Assuming SSP models, these very blue colours translate into an age of less than 10$^{8}$ years, a factor of 10 younger than the youngest cluster populations in the core. This value is supported by assuming tidal interactions between these star forming regions and the centre of the galaxy.
Spatially the young blue populations fall on the bright arms of the H$\alpha$ loop and filament while the older, redder populations are more evenly dispersed around the galaxy. This and the extremely young ages implies the formation mechanism is both separate to the event that formed the star clusters in the centre of the galaxy and is intimately linked to the formation of the H$\alpha$ filaments.
The star formation in the Blue Loop and Southern filament is `clumpy' and there is evidence suggesting that there is a gradient in colours along these `clumps'. This, and the observation of a broad range in colours in the blue population suggest that the event responsible for the star cluster formation is unlikely to have occurred rapidly. However the range in ages of the population are very dependent on the models used and can only be precisely determined with spectroscopy.
We suggest formation mechanisms based on the disruption of the magnetic field supporting the filaments either as a shock from a buoyant radio bubble or by the filaments, fragmenting and falling back into the galaxy after being dragged out below such a bubble. If such a scenario is responsible for the star formation then these outer stellar regions provide another link in the evolutionary cycle feeding the SMBH and regulating the cluster growth. Spectroscopic observations of these regions are needed to confirm that the stellar clusters are really in the LVS and to better determine their ages and kinematics.
The star formation rate in the Blue Loop is $\sim20\mathrm{M}_{\odot}\,\mathrm{yr}^{-1}$. The lifetime of the whole low velocity filamentary system is therefore much greater than 10$^{8}\mathrm{yr}$.
\section{Acknowledgements}
REAC acknowledges STFC for financial support. ACF thanks the Royal Society. REAC would also like to thank Nate Bastian and Nina Hatch for helpful and interesting discussions and the anonymous referee for constructive criticism which has greatly improved this work.
This research has made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration.
\bibliographystyle{mnras}
|
\section{Introduction}
One of the main goals of the present experimental program at the
Tevatron and at the Large Hadron Collider (LHC) is the search for the
Higgs boson(s) in order to elucidate the mechanism of electroweak
symmetry breaking (EWSB). In the Standard Model (SM), the latter is
realized in the most economical way via a single Higgs doublet. This
minimal realization predicts a single neutral Higgs boson, whose mass
can be constrained, from electroweak precision data and the direct
search limit from LEP, to be lighter than $\sim 200$ GeV. However, at
the moment, there is no direct experimental evidence for a neutral
Higgs boson or any other scalar particle like, for example, a charged
boson that can be present in models with a nonminimal Higgs sector.
The LHC is going to explore physics up to the TeV scale in order to
search for the Higgs boson, as well as for any new phenomenon that
would confirm the widespread expectation that the picture of particle
physics in terms of the SM is incomplete. However, new particles with
mass in the TeV range that couple to quarks at the tree level can
modify the predictions for Flavor Changing Neutral Current (FCNC)
processes. Thus any extension of the SM, starting from the simplest we
can think of, a Two-Higgs-Doublet Model (2HDM), needs to face the
problem of avoiding conflicts with the strict limits on FCNC
processes.
Glashow and Weinberg addressed this issue proposing the principle of
Natural Flavor Conservation (NFC) \cite{GW}, which requires that the
matrices of Yukawa couplings to up and down quarks for all the Higgs
fields be diagonal in the basis where the quark mass matrices,
$M^{U,D}$, are diagonal. This implies that, with the exception of models
with vectorlike quarks that mix with the ordinary ones, NFC models do
not have tree-level FCNC couplings. In the 2HDM case, NFC can be realized
imposing the sufficient condition that each of the quark mass matrices
is obtained from a single Higgs field. This can be enforced via a
${\cal Z}_2$ symmetry that acts differently on the two Higgs doublets,
leading to two possibilities usually referred to as type-I (i.e., the
model in which both up and down quarks get their masses from Yukawa
couplings to the same Higgs doublet) and type-II models (where up and
down quarks get their masses from Yukawa couplings to different Higgs
doublets).
A less-restrictive way to suppress FCNC processes, still avoiding
conflict with the experimental bounds, is to consider the criterion of
Minimal Flavor Violation (MFV) \cite{DGIS}, which amounts to assuming
that all the new flavor-changing transitions, including those mediated
at the tree level by electrically neutral particles, are controlled by
the Cabibbo-Kobayashi-Maskawa (CKM) matrix. Thus, the MFV hypothesis
requires that all the flavor-violating interactions of the new
particles present at the TeV scale be linked to the known structure of
the Yukawa couplings.
The enforcement of the MFV hypothesis to the case of multi-Higgs
models has been recently investigated by several groups
\cite{MW,BBR,PT}. In particular, in ref.~\cite{MW} it has been shown,
via group-theoretic arguments, that the MFV hypothesis can be enforced
requiring that all the Higgs Yukawa-coupling matrices be composed from
the pair of matrices $Y^U$ and $Y^D$ that are responsible for the
breaking of the $SU(3)_{Q_{L}} \times SU(3)_{U_{R}} \times
SU(3)_{U_{D}}$ quark flavor symmetry. This requirement restricts the
allowed $SU(3)\times SU(2)\times U(1)$ representations of the Higgs
fields that can couple to the quarks to either be equal to that of the
SM Higgs field, i.e.~$({\bf 1},{\bf 2})_{1/2}$, or transform as $({\bf
8},{\bf 2})_{1/2}$. Examples of the former case, besides the NFC
type I and II models, are the aligned model of ref.~\cite{PT} or the
class of 2HDM presented in ref.~\cite{BGL}. The latter case is quite
different, because the second field does not acquire a vacuum
expectation value (vev) and does not mix with the SM Higgs field. Thus
the scalar spectrum of this model contains a CP-even, color-singlet
Higgs boson (the usual SM one) and three color-octet particles, one
CP-even, one CP-odd and one electrically charged, which are split in
mass proportionally to the SM-Higgs vev \cite{MW}. These colored
scalar particles give rise to an interesting phenomenology for the
LHC, not only because -- if they are not too heavy -- they can be
directly produced, but also because their indirect effects can
influence flavor, electroweak and Higgs physics
\cite{MW,GrW,IKM,BDV,BTZ}.
Models with a second Higgs doublet present a new and interesting
phenomenology, in particular related to the presence of a charged
scalar. In the flavor sector, decays mediated by a weak charged
current are the natural place where effects due to a charged Higgs
boson, $H^+$, can show up. In the electroweak sector the observable
$R_b \equiv \Gamma (\Zbb)/\Gamma(Z \to$ {\em hadrons}) shows a sensitivity to
$H^+$ because of the specific vertex corrections introduced by the
interaction of $H^+$ with the top and bottom quarks. Many studies
(for the most recent see, e.g., refs.~\cite{Gfit,CKMfit,UTfit}) used
various combinations of flavor and electroweak observables to
constrain the parameter space of the type-II 2HDM, which garnered most
of the attention because of its property of having the same
Higgs-sector realization as the Minimal Supersymmetric Standard Model
(MSSM). Other studies \cite{Ida,Moretti,MS} explored the parameter
space of 2HDMs unconstrained by a ${\cal Z}_2$ symmetry, with the
second Higgs doublet still in the $({\bf 1},{\bf 2})_{1/2}$
representation.
The theoretical accuracy of the predictions in the 2HDM with MFV is
not yet at the same level as in the SM. Here we take a first step in
improving this situation, by (re)considering the QCD corrections to
two observables, $R_b $ and ${\rm BR}(\Bsg)$, which allow to set important
constraints on the mass and couplings of the charged scalar. The QCD
corrections can play a relevant role in reducing the error of the
theoretical predictions, a well-known example of this fact being
indeed the radiative decay of the $B$ meson. The present knowledge in
the 2HDMs of the two observables we are considering can be summarized
in this way: in the case of models with a second Higgs doublet in the
$({\bf 1},{\bf 2})_{1/2}$ representation, the complete one-loop
calculation of $\Gamma (\Zbb)$ is available \cite{DGHK}, but (to our
knowledge) no QCD corrections to the charged-scalar contributions are
known. The process $B\to X_{\!s}\, \gamma$ is instead fully known at the
Next-to-Leading Order (NLO) in QCD \cite{CMM,CDGG1,CRS,BG}. The case
with colored scalars in the adjoint representation of $SU(3)$ is less
studied. The one-loop charged-scalar contribution to $\Gamma (\Zbb)$ was
reported in ref.~\cite{GrW} (see also ref.~\cite{HL}) while for the
radiative decay of the B meson only a partial result for the Leading
Order (LO) Wilson coefficients of the magnetic and chromo-magnetic
operators has been presented \cite{MW}.
In this paper we present the QCD corrections to the contribution to
$\Gamma (\Zbb)$ of a charged scalar in either the $({\bf 1},{\bf 2})_{1/2}$
or the $({\bf 8},{\bf 2})_{1/2}$ representation. Concerning $B\to X_{\!s}\, \gamma$, we
compute the ${\cal O}(\alpha_s)$ contribution to the Wilson coefficients
due to a colored charged scalar in the $({\bf 8},{\bf 2})_{1/2}$
representation. This is the missing piece to achieve NLO predictions
for ${\rm BR}(\Bsg)$ for all 2HDMs with MFV. Because of the specific
interactions of the colored scalar with the gluons, the Wilson
coefficients cannot be simply obtained by an appropriate color-factor
rescaling of the known $({\bf 1},{\bf 2})_{1/2}$ result.
The paper is organized as follows: in the next section we discuss the
couplings of the charged Higgs boson in the different realizations of
the 2HDM with MFV. In section 3 we present the results for the
QCD-corrected contribution to $\Gamma (\Zbb)$ due to a charged scalar,
covering both cases of color-singlet and color-octet particle. We show
that, after the inclusion of the QCD correction, the prediction for
$R_b$ is practically insensitive to choice of an $\overline{\rm MS}$ or on-shell
(OS) renormalization scheme for the top mass. The bounds set by $R_b$
on the $tbH^+$ coupling are also shown. Section 4 contains the result
for the NLO Wilson coefficients in the $\Delta B =1$ effective
Hamiltonian (the explicit analytic expressions are presented in the
appendix). The known results for the colorless 2HDMs are recovered,
while the case of a colored charged scalar is fully new. The
restrictions imposed by $B\to X_{\!s}\, \gamma$ on the charged-scalar interaction with
quarks are discussed. Finally, in section 5 we present our
conclusions.
\section{Minimally flavor violating 2HDMs}
In a generic 2HDM it is always possible to rotate the two Higgs fields
to a basis in which only one of them, which we denote as $\Phi_1$,
gets a vev~\cite{habdav}. In this basis, we write the Yukawa
interactions of the Higgs fields with the quarks as
\begin{equation}
\label{lagYuk}
-{\cal L}_Y ~=~
\bar{q}_L\, \widetilde{\Phi}_1\, Y^U u_R ~+~ \bar{q}_L\, \Phi_1\, Y^D d_R
~+~\bar{q}_L \,\widetilde{\Phi}^{(a)}_2\, T_R^{(a)}\,\bar{Y}^U u_R ~+~
\bar{q}_L \,\Phi_2^{(a)}\, T_R^{(a)}\, \bar{Y}^D d_R~~+~{\rm h.c.}~,
\end{equation}
where $\widetilde{\Phi}_i \equiv i\sigma_2\Phi_i^*$, and the Yukawa
couplings $Y^{U,D}$ are $3\!\times\!3$ matrices in flavor space such
that $M^{U,D} = Y^{U,D}\,\langle \Phi_1^0\rangle$. The possibility of
a colored second Higgs doublet is encoded in the matrices $T_R^{(a)}$
that act on the quark fields. In the usual colorless 2HDM $T_R$ is
equal to the identity matrix in color space. On the other hand, for a
colored Higgs doublet in the adjoint representation of $SU(3)$
$T_R^a=T_F^a\: (a=1,8)$, the matrices of the fundamental
representation. The MFV condition amounts to requiring that the
Yukawa coupling matrices of the second doublet, $\bar Y^{U,D}$, be
composed of combinations of the matrices $Y^{U,D}$, and transform
under the $SU(3)_{Q_{L}} \times SU(3)_{U_{R}} \times SU(3)_{U_{D}}$
quark flavor symmetry in the same way as $Y^{U,D}$ themselves. We can
decompose the matrices $\bar Y^{U,D}$ as
\begin{equation}
\label{MFVdeco}
\bar Y^U = A_u \,\left( 1 + \epsilon_u\,Y^U Y^{U\,\dagger}
+ \ldots\right)\,Y^U~,~~~~~
\bar Y^D = A_d \,\left( 1 + \epsilon_d\,Y^U Y^{U\,\dagger}
+ \ldots\right)\,Y^D~,
\end{equation}
where in principle $A_{u,d}$ and $\epsilon_{u,d}$ are arbitrary
complex coefficients. The ellipses in eq.~(\ref{MFVdeco}) denote terms
involving powers of $Y^DY^{D\,\dagger}$ as well as terms involving
higher powers of $Y^UY^{U\,\dagger}$. In the following, we will assume
that the only significant deviations from proportionality between
$\bar Y^{U,D}$ and $Y^{U,D}$ are controlled by the Yukawa coupling of
the top quark, and that terms involving higher powers of the Yukawa
matrices are suppressed (e.g., because they are generated at higher
loops). If we further require that there are no new sources of CP
violation apart from the complex phase in the CKM matrix, the
coefficients $A_{u,d}$ and $\epsilon_{u,d}$ must be real. Finally,
the case $\epsilon_u=\epsilon_d=0$ corresponds to the NFC situation in
which the Yukawa matrices of both Higgs doublets are aligned in flavor
space.
The processes $Z \to b \bar{b}$ and $B\to X_{\!s}\, \gamma$ that we will consider in sections 3
and 4 involve loops with a charged Higgs boson and a top quark. Under
the assumptions implicit in eq.~(\ref{MFVdeco}), the interaction
between the quarks and $H^+$ is controlled by the Lagrangian
\begin{equation}
\label{lagH}
{\cal L}_{H^+}= - \frac{g}{\sqrt{2\,} m_\smallw}\,
\sum_{i,j=1}^3\,\bar u_i \,T_R^{(a)}
\left( A^i_u \,m_{u_i}\,\frac{1-\gamma_5}{2}-
A^i_d\, m_{d_j}\,\frac{1+\gamma_5}{2}\right)\, V_{ij}\,d_j\,
H^+_{(a)} +{\rm h.c.}~,
\end{equation}
where $g$ is the $SU(2)$ coupling constant, $i,j$ are generation
indices, $m_{u,d}$ are quark masses, $V$ is the CKM matrix. The
family-dependent couplings $A_{u,d}^i$ read
\begin{equation}
\label{shift}
A_{u,d}^i = A_{u,d}\,\left(1 +
\epsilon_{u,d}\,\frac{m_t^2}{v^2}\,\delta_{i3}\right)~,
\end{equation}
where $v = \langle \Phi_1^0 \rangle$. It appears from
eq.~(\ref{shift}) that, when we neglect the masses of the light
quarks, the effect on the charged-Higgs couplings arising from the
$Y^UY^{U\,\dagger}$ terms in eq.~(\ref{MFVdeco}) is limited to a shift
in the couplings involving the top quark. Since those are the only
couplings that enter our computations, in sections 3 and 4 we will
drop the family index from $A_{u,d}^i$ without ambiguity.
The term $\epsilon_d Y^UY^{U\,\dagger}$ entering the expression for
$\bar Y^D$ in eq.~(\ref{MFVdeco}) also induces a flavor-changing
interaction with the down quarks for the neutral component of the
second Higgs doublet. However, this interaction does not affect the
computation of $\Gamma (\Zbb)$, and its contribution to $B\to X_{\!s}\, \gamma$ is negligible
with respect to the charged-Higgs contribution as long as
$\epsilon_d\,A_d/A_u \ll (v/m_b)^2$. Other FCNC processes such as
$B\bar B$ mixing would put bounds on the combination
$\epsilon_d\,A_d$, but this will not be relevant to the discussion
that follows.
In the notation of eq.~(\ref{lagH}), the type-I and type-II models are
specified by $T_R^a$ equal to the identity matrix in color space, and
by the real (and family-universal) coefficients
\begin{eqnarray}
A_u^i = A_d^i &=& 1/\tan \beta ~~~~~~ \mbox{(type I)}, \\
A_u^i=-1/A_d^i &=& 1/\tan \beta ~~~~~~ \mbox{(type II)},
\end{eqnarray}
where $\tan \beta$ is the ratio of the vevs of the two Higgs doublets
in the basis where each of the quark mass matrices is obtained from a
single Higgs field.
According to our discussion, the MFV hypothesis includes two other
possibilities, namely color-singlet and color-octet Higgs doublet that
couple to the quarks with arbitrary coefficients\footnote{Even models
with generic Yukawa matrices not satisfying the MFV hypothesis show a
structure of couplings with arbitrary coefficients $A_{u,d}^i$. To
be phenomenologically viable, the dangerous FCNC effects should be
sufficiently suppressed via some specific assumption like, e.g., a
specific texture of the Yukawa matrices \cite{CS}.} $A_{u,d}^i$. We
are going to refer to the first category (singlet) as type-III model,
while the second possibility (octet) will be called type-C model.
\section{Charged-Higgs contribution to $R_b$ including QCD
corrections}
We begin by discussing the radiatively corrected partial decay width
of the $Z$ boson in a quark-antiquark pair in a model with an
additional Higgs doublet. We write it as
\begin{equation}
\Gamma(Z \to q \bar{q}) = N_c\, \frac{G_\mu}{\sqrt2}\, \frac{m_\smallz^3}{3\, \pi}
\,\left[ (\bar{g}^{q}_L)^2 + (\bar{g}^{q}_R)^2 \right] K_q~,
\label{eq:0}
\end{equation}
where $N_c$ is the color factor ($N_c=3$), $\bar{g}^{q}_{(L,R)}$ are
the left-handed and right-handed $Zq\bar q$ couplings, written in terms
of the radiative parameter $\rho_q$ and the radiatively corrected sine
of the Weinberg angle $\bar{s}_W^q$ as ($T_3^q$ is the third component
of the weak isospin, $Q_q$ is the electric charge in unit $e$)
\begin{eqnarray}
\bar{g}^{q}_L &=& \sqrt{\rho_q} \,\left[ T_3^q
- Q_q\, (\bar{s}_W^q)^2 \right]~,\nonumber \\
\bar{g}^{q}_R &=& - \sqrt{\rho_q}\, Q_q (\bar{s}_W^q)^2~,
\end{eqnarray}
while the factor $K_q$ contains the QCD, QED and quark-mass
corrections. The latter has been computed up to ${\cal O}(\alpha_s^3)$ in
ref.~\cite{CKK}, and at the lowest order it reads:
\begin{equation}
K_q = 1 + C_F \frac{3\alpha_s}{4 \pi} + Q_q^2\, \frac{3\alpha}{4 \pi} -
\frac34 \frac{\mu_q}{(\bar{g}^{q}_L)^2 + (\bar{g}^{q}_R)^2} + {\cal O}(\alpha_s^2)~,
\end{equation}
where $\mu_q = m_q^2/m_\smallz^2$ and $C_F =4/3$.
We assume that the oblique corrections due to the second doublet are
negligible, as happens when the spectrum of the additional states is
approximately custodially symmetric. Then the effect of the second
Higgs doublet is concentrated in the vertex corrections to
$\Gamma (\Zbb)$. Thus, defining
\begin{eqnarray}
\rho_q &=& \rho_q^{SM} + \delta \rho_q~, \\
(\bar{s}_W^q)^2 &=& (\bar{s}_{W}^q)^2_{SM} + \delta (\bar{s}_W^q)^2~,
\end{eqnarray}
we have $\delta \rho_{(q \neq b)} = \delta (\bar{s}_W^{(q \neq b)})^2
=0$. In the limit of neglecting the mass of the $Z$ boson with respect
to the masses of the top quark and the charged Higgs boson we find
\begin{eqnarray}
\delta \rho_b \hspace{-0.15cm} &=&
\frac1{T_3^b}\,\frac{\alpha}{4 \pi s^2_W} \,C^1_R\,
\left[ \left( \frac{|A_u|\,\hat{m}_t}{\sqrt{2} m_\smallw}\right)^2 +
\left(\frac{|A_d| \,\hat{m}_b}{\sqrt{2} m_\smallw}\right)^2
\right] \left[ \,f_1 (t_h) + \frac{\alpha_s}{4 \pi}
\left(C_F \, f_2 (t_h) + C_R^2 \, f_3 (t_h)
\right) \right]\,,~~~ \label{eq:1}\\\nonumber\\
\delta (\bar{s}_W^b)^2 \hspace{-0.15cm} &=& -\frac12 \delta \rho_b s^2_W +
\frac1{2\,Q_b} \, \frac{\alpha}{4 \pi s^2_W} \, C^1_R
\left(\frac{|A_d| \,\hat{m}_b}{\sqrt{2} m_\smallw}\right)^2
\left[ f_1(t_h) + \frac{\alpha_s}{4 \pi}
\left(C_F \,f_2 (t_h) + C_R^2 \, f_3 (t_h)
\right) \right]~,
\label{eq:2}
\end{eqnarray}
where $C^1_R=1, C^2_R = 0 \:[C^1_R=C_F, C^2_R = C_A=N_c]$ for Higgs
fields in the $({\bf 1},{\bf 2})_{1/2}\: [({\bf 8},{\bf 2})_{1/2}]$
representation, and we omit an overall factor $|V_{tb}|^2 \approx
1$. In eqs.~(\ref{eq:1}) and (\ref{eq:2}) $t_h= \hat{m}_t^2/m_\smallh^2$, where
$\hat{m}_q$ is the $\overline{\rm MS}$ quark mass at the scale $\mu$ and $m_\smallh$ is
the OS $H^+$ mass. The explicit expressions for the functions
$f_i(x)$ are
\begin{eqnarray}
f_1(x) & = & \frac{x}{x-1}-\frac{x \,\ln x}{(x-1)^2}~, \\
f_2(x) & =&
-\frac{6 \,x\left(x-2 \right)}{(x-1)^2} {\rm Li}_2 \left(1-\frac1x\right)+
\frac{ x (-27 +11 x)}{(x-1)^2}+\frac{x (25-9 x) \ln x}{ (x-1)^3}\nonumber \\
&+&
\left( \frac{6x(3-x)}{(x-1)^2} - \frac{12 x \ln x}{(x-1)^3} \right)
\ln \frac{\hat{m}_t^2}{\mu^2} -3 f_1(x)~,\\
f_3(x) & = & \frac{3
x }{ (x-1)} {\rm Li}_2\left(1-\frac1x\right)
+\frac{3 x \left(1-2 x + x^2+ \ln^2 x\right)}{ (x-1)^3}-
\frac{6 x \ln x}{(x-1)^2}~,
\end{eqnarray}
where the last term ($-3 f_1$) in the function $f_2$ is introduced to
avoid double counting due to the correction factor $K_q$ in
eq.~(\ref{eq:0}), and in the $tbH^+$ coupling we have also kept the
contribution proportional to the bottom mass\footnote{Terms
proportional to $\hat{m}_b$, relevant only for very large values of $A_d$,
can also arise from vertices with neutral scalars.}.
The one-loop terms in $\delta \rho_b$ and $\delta (\bar{s}_W^b)^2$
agree with the results\footnote{In ref.~\cite{HL} there is a misprint
in the overall normalization of the $\delta g^{L,R}$ couplings.} of
refs.~\cite{DGHK,HL}. The two-loop terms were obtained following the
lines of the analogous SM calculation that was performed by several
groups, via different methods, in the early nineties \cite{Zbbme}. The
SM correction can be actually obtained by considering the SM
Lagrangian in the limit of vanishing gauge coupling constants, the
so-called gaugeless limit of the SM \cite{BBCCV}. In this limit the
gauge bosons play the role of external sources, and the propagating
fields are those of a Yukawa theory with massless Goldstone
bosons. Indeed, in the limit $m_\smallh \to 0$, the ${\cal O}(\alpha \alpha_s
\hat{m}_t^2/m_\smallw^2)$ corrections in eq.~(\ref{eq:1}) agree with the known SM
result.
To express the corrections in terms of the OS top mass $m_t$, we
must expand the $\overline{\rm MS}$ top mass entering the one-loop part as
$\hat{m}_t = m_t + \delta m_t$, with
\begin{equation}
\delta m_t = \frac{\alpha_s}{4 \pi} \,C_F \left( 3 \ln
\frac{m_t^2}{\mu^2} -4 \right) \, m_t~.
\end{equation}
For the terms proportional to $|A_u|^2$ and $|A_d|^2$ in
eqs.~(\ref{eq:1}) and (\ref{eq:2}), this amounts to making the
substitution $\hat{m}_t \to m_t$ and replacing the function $f_2$ with the
OS counterparts $f_2^u$ and $f_2^d$, respectively:
\begin{equation}
f_2^{u}(x) ~=~
f_2(x) ~+~ \frac{8\pi}{\alpha_s\,C_F}\,\frac{\delta m_t}{m_t}\,
\left[f_1(x) + x\,\frac{\partial f_1(x)}{\partial x}\right]~,~~~~~~
f_2^{d}(x) ~=~
f_2(x) ~+~ \frac{8\pi}{\alpha_s\,C_F}\,\frac{\delta m_t}{m_t}\,x\,\frac{\partial f_1(x)}{\partial x}~.
\end{equation}
The explicit dependence on the renormalization scale cancels out in
the function $f^u_2$:
\begin{equation}
f_2^{u}(x)
~=~
-\frac{6 \,x\left(x-2 \right)}{(x-1)^2} {\rm Li}_2 \left(1-\frac1x\right)+
\frac{ 3 x }{(x-1)}-\frac{9 x \ln x}{ (x-1)^2}-3 f_1(x)~,
\end{equation}
whereas $f^d_2$ has a residual dependence on $\mu$ which is
compensated for by the implicit scale dependence of the $\hat m_b^2$
entering the one-loop parts of eqs.~(\ref{eq:1}) and
(\ref{eq:2}). Using the OS bottom mass $m_b$ would remove this
residual scale dependence, but it would introduce large logarithms of
the ratio $m_t/m_b$ in the two-loop part of the corrections.
\begin{figure}[t]
\begin{center}
\epsfig{figure=dqcd_vs_mH.eps,width=13cm}
\end{center}
\vspace{-2mm}
\caption{the ratio of two-loop to one-loop charged-Higgs contributions
to the $Zb\bar b$ vertex as a function of $m_\smallh$, with the top mass
expressed in the $\overline{\rm MS}$ (solid lines) or OS (dashed lines)
renormalization scheme. The upper (red) curves are for the model
with color-octet Higgs (type C), while the lower (blue) curves are
for the models with color-singlet Higgs (types I--III).}
\label{fig1}
\end{figure}
In fig.~\ref{fig1} we plot, as a function of $m_\smallh$, the ratio
$\delta^{QCD}$ between the two-loop and one-loop contributions in the
terms proportional to $|A_u|^2$ in eq.~(\ref{eq:1}), for the two cases
of color-singlet (lower set of lines) and color-octet (upper set of
lines) charged Higgs boson. For each case, we show $\delta^{QCD}$ as
obtained using either the central value of the physical (OS) top mass,
$m_t = 173.1$ GeV \cite{top} (dashed lines), or the corresponding
$\overline{\rm MS}$ value $\hat{m}_t(m_t)= 163.5$ GeV (solid lines), with the
appropriate formulae for the two-loop function $f_2$. It can be seen
from the figure that, for models of types I--III (i.e.~with
color-singlet charged Higgs) the two-loop corrections are always
negative, and they are substantially larger when the OS top mass is
used in the one-loop part than when the $\overline{\rm MS}$ mass is used. On the
other hand, for the model of type C (with color-octet charged Higgs)
there is an overall upward shift in the two-loop correction due to the
additional function $f_3$ in eq.~(\ref{eq:1}), with the result that
the sign and relative size of the corrections in the OS and $\overline{\rm MS}$
cases depend on the Higgs mass. For low values $m_\smallh \approx 150$ GeV,
the two-loop correction approaches zero if the OS top mass is used,
and it is positive and relatively large if the $\overline{\rm MS}$ mass is
used. Conversely, for larger values $m_\smallh \approx 400$ GeV the two-loop
correction approaches zero if the $\overline{\rm MS}$ top mass is used, and it is
negative and relatively large if the OS mass is used. As a result, we
will see that for 2HDMs of types I--III a reliable approximation of
the two-loop result for $R_b$ could be obtained by using the one-loop
result expressed in terms of the $\overline{\rm MS}$ top mass. On the other hand,
a precise determination of $R_b$ in the 2HDM of type C requires the
inclusion of the two-loop part of the $Zb\bar b$ vertex correction.
From eqs.~(\ref{eq:0})--(\ref{eq:2}) we can construct the observable
$R_b $, which can be written as
\begin{equation}
\label{defRb}
\frac1{R_b} = 1 + \frac{\sum_{(q \neq b)}
\left[ (\bar{g}^{q}_L)^2 + (\bar{g}^{q}_R)^2 \right] K_q}
{ \left[ (\bar{g}^{b}_L)^2 + (\bar{g}^{b}_R)^2 \right] K_b}
\equiv 1+ \frac{S_b}{s_b} C_{b}
\end{equation}
where
\begin{equation}
\label{defSb}
S_b = \sum_{(q \neq b)} s_q~;~~~~~~~
s_q = \left[ (\bar{g}^{q}_L)^2 + (\bar{g}^{q}_R)^2 \right]
\left( 1 + Q_q^2\, \frac{3\alpha}{4 \pi} \right)~.
\end{equation}
Using the results of ref.~\cite{LEPEWWG} we find
$S_b=0.6607$. Concerning the SM part of $s_b$, to avoid relying
indirectly on the measured value of $R_b$, we follow
ref.~\cite{CKMfit} and compute it using the values $\rho_b^{SM} =
0.99426$ \cite{LEPEWWG} and $(\bar{s}_W^b)^2_{SM} = (1.0063)\times
0.23153=0.23299$, the latter obtained from the measured value of
$\sin^2 \theta^{lept}_{eff}=0.23153 \pm 0.00016$ corrected for the
top-induced contributions specific to the $Zb\bar{b}$ vertex\footnote{
In ref.~\cite{CKMfit} the correcting factor 1.0063 was not
introduced.}. Finally the factor $C_{b}$ that includes QCD and mass
corrections is obtained from ref.~\cite{CKK}. We find, for $\alpha_s =
0.118$, $C_b = 1.0086$.
With the values specified above for the various parameters entering
eqs.~(\ref{defRb}) and (\ref{defSb}) we find a SM prediction
$R_b^{\rm{\scriptscriptstyle SM}} = 0.21580$, nearly $1\sigma$ below
the measured value $R_b^{\rm exp} = 0.21629 \pm 0.00066$
\cite{LEPEWWG}. Since the charged-Higgs contributions to
$\bar{g}_{L}^b$ and $\bar{g}_{R}^b$ have the effect of further
lowering $R_b$, stringent bounds can be imposed on the parameters
$m_\smallh$ and $A_u$ by the requirement that the predicted value of $R_b$
in a 2HDM be not too far from $R_b^{\rm exp}$. On the other hand,
$R_b$ has little sensitivity on $A_d$, because the terms in
eqs.~(\ref{eq:1}) and (\ref{eq:2}) controlled by the latter are
suppressed by $\hat m_b$. Therefore, for the models of types III and C
in which $A_d$ is a free parameter we will simplify our discussion by
setting $A_d=0$. In the 2HDMs of type I and II the parameter $A_d$ is
related to $A_u$ and cannot be set independently to zero. However, due
to the strong suppression of the contributions controlled by $A_d$, in
most of the parameter space the predictions of $R_b$ obtained in those
two models do not differ significantly from the predictions obtained
in the type-III 2HDM with $A_d=0$. More specifically, the predictions
of the type-I 2HDM, in which $A_d=A_u$, are virtually
indistinguishable from those of the type-III 2HDM with $A_d=0$ for all
the values of $A_u$ consistent with the measured value of $R_b^{\rm
exp}$. In the 2HDM of type II, on the other hand, $A_d=-1/A_u$, and
the predictions of $R_b$ differ from the ones obtained in the type-III
2HDM with $A_d=0$ only for very small values of $A_u$.
\begin{figure}[p]
\begin{center}
\epsfig{figure=Rb_vs_Au-2H.eps,width=13cm}
\end{center}
\vspace{-2mm}
\caption{$R_b$ as a function of $|A_u|$ in the 2HDM with color-singlet
Higgs, for $A_d=0$ and two different values of $m_\smallh$. The measured
value $R_b^{\rm exp} = 0.21629$ and the values $1\sigma$ and
$2\sigma$ below it are displayed as horizontal lines. For the
meaning of the different curves see the text.}
\label{fig2}
\end{figure}
\begin{figure}[p]
\begin{center}
\epsfig{figure=Rb_vs_Au-MW.eps,width=13cm}
\end{center}
\vspace{-2mm}
\caption{same as figure \ref{fig2} in the 2HDM with color-octet
Higgs.}
\label{fig3}
\end{figure}
Figures \ref{fig2} and \ref{fig3} show our determination of $R_b$ in
the 2HDMs of type III and C, respectively, as a function of
$|A_u|$. In each plot we show two sets of curves for the charged-Higgs
mass values $m_\smallh = 100$ GeV and $m_\smallh = 400$ GeV. In each set, the
dashed (solid) curve represents the one-loop (two-loop) result
expressed in terms of the $\overline{\rm MS}$ top mass, while the dotted
(dot-dashed) curve represents the one-loop (two-loop) result expressed
in terms of the physical top mass. We also show in each plot the
measured value $R_b^{\rm exp}$ (solid horizontal line) and the values
$1\sigma$ and $2\sigma$ below (dashed horizontal lines). It can be
seen that, in both plots, the curves corresponding to the two-loop
results (with the top mass renormalized either in the $\overline{\rm MS}$ or in
the OS scheme) are practically overlapped. The location of the curves
corresponding to the one-loop results reflects the behavior that could
already be inferred from fig.~\ref{fig1}: in the 2HDM of type III the
one-loop result computed in terms of the $\overline{\rm MS}$ top mass is very
close to the two-loop result, while the one-loop result computed in
terms of the physical top mass can differ significantly. On the other
hand, in the 2HDM of type C the quality of the one-loop approximation
depends on the charged Higgs mass. At low values of $m_\smallh$, using the
physical top mass in the one-loop result gives a much better
approximation to the two-loop result than using the $\overline{\rm MS}$ top mass,
while the situation is reversed at large values of $m_\smallh$. Therefore,
only the use of the two-loop results guarantees a precise
determination of $R_b$ for all the values of $m_\smallh$.
From figures \ref{fig2} and \ref{fig3} it is also possible to
determine the values of $|A_u|$ that are disfavored by the comparison
between $R_b^{\rm exp}$ and the corresponding theoretical prediction.
In the case of the type-III 2HDM, the two-loop curves cross the
$2\sigma$ horizontal line at $|A_u| = 0.78$ for $m_\smallh = 100$ GeV and at
$|A_u| = 1.35$ for $m_\smallh = 400$ GeV. In the case of the type-C 2HDM,
they cross it at $|A_u| = 0.62$ for $m_\smallh = 100$ GeV and at $|A_u| =
1.10$ for $m_\smallh = 400$ GeV. We checked that the crossing points for
intermediate values of $m_\smallh$ can be determined by linear interpolation
of the values given above.
The predictions of $R_b$ for the type-II 2HDM (where $A_d=-1/A_u =
-\tan\beta$) are virtually indistinguishable from those presented in
figure \ref{fig2} for the type-III 2HDM as soon as $|A_u|>0.1$. The
upper bounds on $|A_u|$ discussed above translate, for the type-II
2HDM, into $|A_d| > 1.28$ for $m_\smallh = 100$ GeV and $|A_d| > 0.74$ for
$m_\smallh = 400$ GeV. On the other hand, when $|A_u|$ tends to zero the
predictions of $R_b$ in the type-II 2HDM decrease quickly, and get two
standard deviations below $R_b^{\rm exp}$ for $|A_u| \approx 0.01$,
corresponding to $|A_d| \approx 100$. As we will see in the next
section, the bounds on $A_d$ coming from the process $B\to X_{\!s}\, \gamma$ can be much
stronger than that, but they do not apply to the type-II 2HDM.
\section{Charged-Higgs contribution to $B\to X_{\!s}\, \gamma$ at the NLO}
The branching ratio for $B\to X_{\!s}\, \gamma$ is fully known at the NLO for a 2HDM of
types I--III. To cover also the case of a type-C 2HDM, the only
missing ingredient is the determination of the colored-scalar ${\cal
O}(\alpha_s)$ contribution to the Wilson coefficients. We compute it
following the analogous computation for the type I--II 2HDM presented
in ref.~\cite{CDGG1}.
In the operator basis defined in ref.~\cite{CDGG1} we write the Wilson
coefficients at the scale $\mu_W$, where the ``full" theory is matched
to an effective theory with five quark flavors, as
\begin{equation}
C^{}_i(\mu_W) = C^{(0)}_i(\mu_W) + \delta C_{i}^{(0)}(\mu_W)
+ \frac{\alpha_s(\mu_W)}{4 \pi} \left[ C^{(1)}_i(\mu_W) +
\delta C_{i}^{(1)}(\mu_W) \right]~,
\label{wc1}
\end{equation}
where $C^{(k)}_i(\mu_W)$ represents the SM contribution ($k=0,1$) while
$ \delta C_{i}^{(k)}(\mu_W)$ represents the charged-Higgs
contribution. At the LO, the latter is given by
\begin{eqnarray}
\delta C_{i}^{(0)}(\mu_W) &= &0 ~~~~~i=1,...,6~,\\
\delta C_{7}^{(0)}(\mu_W) &= & C^1_R \left( \frac{|A_u|^2}3
F_{7}^{(1)}(y) - A_d A_u^* F_{7}^{(2)}(y) \right)~,
\label{eqc7} \\
\delta C_{8}^{(0)}(\mu_W) &= & C^1_R \left( \frac{|A_u|^2}3
F_{8}^{(1)}(y) -A_d A_u^* F_{8}^{(2)}(y) \right) + C^2_R \left(
|A_u|^2 F_{8}^{(3)}(y) + A_d A_u^* F_{8}^{(4)}(y)\right)~,
\label{eqc8}
\end{eqnarray}
where
\begin{equation}
F_7^{(1)}(y)=\frac{y(7-5y-8y^2)}{24(y-1)^3}+\frac{y^2(3y-2)}{4(y-1)^4}\ln y,
~~~~~~~~F_7^{(2)}(y)= \frac{y(3-5y)}{12(y-1)^2}+\frac{y(3y-2)}{6(y-1)^3}\ln y,
\end{equation}
\begin{equation}
F_8^{(1)}(y)= \frac{y(2+5y-y^2)}{8 (y-1)^3}-\frac{3y^2}{4(y-1)^4}\ln y,
~~~~~~~~F_8^{(2)}(y)=\frac{y(3-y)}{4(y-1)^2}-\frac{y}{2(y-1)^3}\ln y,
\end{equation}
\begin{equation}
F_8^{(3)}(y)=\frac{y(1+y)}{16(y-1)^2}-\frac{y^2}{8(y-1)^3}\ln y,
~~~~~~~~F_8^{(4)}(y)=-\frac{y}{4(y-1)}+\frac{y}{4(y-1)^2}\ln y,
\end{equation}
with
\begin{equation}
y=\frac{\hat m_t^2(\mu_W )}{m_\smallh^2}~,
\end{equation}
expressed in terms of the NLO top-quark running mass at the scale
$\mu_W$ and of the OS charged-Higgs mass. Again, $C^1_R=1, C^2_R = 0
\:[C^1_R=C_F, C^2_R = C_A =N_c]$ for Higgs fields in the $({\bf
1},{\bf 2})_{1/2}\: [({\bf 8},{\bf 2})_{1/2}]$ representation.
At the NLO, the charged-Higgs contributions to the Wilson coefficients
are
\begin{eqnarray}
\delta C_{i}^{(1)}(\mu_W) & = & 0 ~~~~~i=1,2,3,5,6 ~,\\
\delta C_{4}^{(1)}(\mu_W) & = & E^H(y)~,\\
\delta C_7^{(1)}(\mu_W) &=& G_7^H(y) + \Delta_7^H(y) \ln\frac{\mu_W^2}{m_\smallh^2}~,
\label{c7NLO} \\
\delta C_8^{(1)}(\mu_W) &=& G_8^H(y) + \Delta_8^H(y) \ln\frac{\mu_W^2}{m_\smallh^2}~.
\label{c8NLO}
\end{eqnarray}
The expressions for $G^H_{7,8}, \:\Delta^H_{7,8}$ and $E^H$ are rather
long and they are reported in the appendix. As expected, the $\mu_W$
dependence in $\delta C_{7,8}$ cancels out at ${\cal O}( \alpha_s)$ because
the functions $\Delta^H_{7,8}$ entering eqs.~(\ref{c7NLO}) and
(\ref{c8NLO}) satisfy the relation
\begin{equation}
\Delta^H_i=\gamma_0^m y \frac{\partial\, \delta C_i^{(0)}}{\partial
y}+\frac{1}{2} \sum_{j=1}^8\gamma_{ji}^{(0)eff} \delta C_j^{(0)}~,
\label{canc}
\end{equation}
where $\gamma_0^m = 8$ is the LO anomalous dimension of the top mass,
while $\gamma_{ji}^{(0)eff}$ is the matrix of LO anomalous dimensions
of the Wilson coefficients, whose entries can be found in eq.~(8) of
ref.~\cite{CMM}.
\begin{figure}[t]
\begin{center}
\epsfig{figure=dqcdBR_vs_mH_AdAu.eps,width=13cm}
\end{center}
\vspace{-2mm}
\caption{the ratio of NLO to LO charged-Higgs contributions to the
Wilson coefficients $C_7$ (solid line) and $C_8$ (dashed line) as a
function of $m_\smallh$, with $A_d = A_u$, for the model with color-octet
Higgs (type C) or color-singlet Higgs (type III).}
\label{fig3bis}
\end{figure}
In fig.~\ref{fig3bis} we show the ratio between the NLO and LO
charged-Higgs contributions to the Wilson coefficients $C_{7,8}$ as a
function of $m_\smallh$, for both the type-C (color-octet) and type-III
(color-singlet) cases, with the particular choice $A_d=A_u$ (in the
color-singlet case this coincides with the type-I 2HDM). For the
type-C 2HDM, the NLO corrections can reach up to $\sim 20\%$ of the LO
contribution at small $m_\smallh$, and they decrease as the charged-Higgs
mass increases, eventually crossing zero. In contrast, for the
type-III 2HDM the two-loop corrections to the Wilson coefficients are
always different from zero, and have opposite sign with respect to the
LO contributions. We checked that the behavior of the NLO corrections
is qualitatively similar to the one described above even when we allow
$A_d$ to take on values different from $A_u$.
The calculation of ${\rm BR}(\Bsg)$ is performed using a modified version of
the fortran code {\tt SusyBSG} \cite{SBSG}. The code provides a NLO
evaluation of the branching ratio in the MSSM with MFV, including the
full two-loop gluino contributions to the Wilson coefficients
\cite{DGS}. The current public version (1.3) includes also the options
of evaluating the branching ratio in the SM, in the type-II 2HDM and
in the MSSM with two-loop gluino contributions computed in the
effective Lagrangian approximation. We enlarged the 2HDM option to
include also the type-I, type-III and type-C models, thus covering all
four types of 2HDM compatible with MFV.\footnote{A public version of
{\tt SusyBSG} with this new feature will be released soon.} The
relation between the Wilson coefficients and ${\rm BR}(\Bsg)$ is computed at
NLO along the lines of ref.~\cite{GM}, but the free renormalization
scales entering the NLO calculation are adjusted in such a way as to
mimic the Next-to-Next-to-Leading Order (NNLO) contributions presented
in ref.~\cite{Mal}. When the SM input parameters are set to the
partially outdated values used in ref.~\cite{Mal}, {\tt SusyBSG} gives
a SM prediction for ${\rm BR}(\Bsg)$ of $3.15 \times 10^{-4}$, in full
agreement with the NNLO result of that paper. Very good numerical
agreement is also found with the results of the partial NNLO
implementation of the type-II 2HDM in ref.~\cite{Mal}, which combines
NNLO anomalous dimensions and matrix elements with NLO Wilson
coefficients. We take into account a recent update \cite{GG} in the
calculation of the normalization factor for the branching ratio as
well as the latest central value of the top mass \cite{top}, which
results in a modest enhancement of the SM prediction for ${\rm BR}(\Bsg)$ to
$3.28 \times 10^{-4}$.
\begin{figure}[t]
\begin{center}
\epsfig{figure=BR_vs_Ad_same.eps,width=8.4cm}
\epsfig{figure=BR_vs_Ad_oppo.eps,width=8.4cm}
\end{center}
\vspace{-1cm}
\caption{${\rm BR}(\Bsg)$ as a function of $A_d$ in the type-III (solid line)
and type-C (dashed line) models for $A_u=0.3$ (left panel) and
$A_u=-0.3$ (right panel) and two different values of $m_\smallh$. The
horizontal dashed lines specify the $2 \sigma$ interval around the
experimental value of ${\rm BR}(\Bsg)$.}
\label{fig5}
\end{figure}
In the 2HDMs of types I and II, the requirement of consistency between
the theoretical prediction and the measured value of ${\rm BR}(\Bsg)$ allows
us to set bounds on the parameters $m_\smallh$ and $\tan\beta$, the latter
determining both Higgs-quark couplings $A_u$ and $A_d$. More
specifically, in the type-I 2HDM the charged-Higgs contribution to the
Wilson coefficients $C_{7,8}$ scales like $1/\tan\beta^2$, therefore
it is possible to derive, for each given value of $m_\smallh$, a lower bound
on $\tan\beta$ (i.e., an upper bound on $A_u=A_d$) which is however
much less stringent than the corresponding bound derived from
$R_b$. In the type-II 2HDM there is a $\tan\beta$-independent
contribution to the Wilson coefficient, which allows to set an
absolute lower bound on $m_\smallh$. These bounds have been extensively
discussed in the literature (see, e.g.,
refs.~\cite{Gfit,CKMfit,UTfit}) and we will not further consider them
here.
In the models of type III and C, on the other hand, the parameters
$A_u$ and $A_d$ are unrelated to each other. As can be seen in
eqs.~(\ref{eqc7}) and (\ref{eqc8}), the charged-Higgs contributions to
the Wilson coefficients $C_{7,8}$ include two terms controlled by
$|A_u|^2$ and $A_d A_u^*$, respectively. The bounds on $A_u$ derived
in the previous section tell us that the former term cannot be too
large, while the latter can be significant for large values of
$A_d$. Furthermore, its effect on the branching ratio depends on the
relative sign between $A_u$ and $A_d$.
In fig.~\ref{fig5} we show ${\rm BR}(\Bsg)$ as a function of $A_d$, for both
the type-III (solid lines) and type-C (dashed lines) 2HDM, and for the
representative choices $A_u = \pm 0.3$ (the latter are allowed by
$R_b$, as can be seen in figs.~\ref{fig2} and \ref{fig3}). The left
panel displays the case of same sign between $A_u$ and $A_d$, while
the right panel shows the case of opposite sign. Each plot contains
two sets of curves corresponding to $m_\smallh = $ 100 GeV and $m_\smallh = 400$
GeV, respectively. The horizontal dashed lines mark the 95\% C.L.~band
around the experimental value ${\rm BR}(\Bsg) = (3.52 \pm 0.25)\times
10^{-4}$ \cite{HFAG}. The band also includes, added in quadrature, the
theoretical error on the 2HDM prediction (we conservatively estimate
this error as 10\% of the SM prediction for the branching ratio).
From fig.~\ref{fig5} it is clear that, unless $|A_u|$ is extremely
small, the process $B\to X_{\!s}\, \gamma$ sets stringent limits on $A_d$. Focusing on
the case of color-singlet Higgs we see that, for $A_u = 0.3$, the
values of $A_d$ that allow for a branching ratio inside the 95\%
C.L.~band are $A_d \leq 0.9$ for $m_\smallh = 100$ GeV and $A_d \leq
2.5$ for $m_\smallh= 400$ GeV; for $A_u = -0.3$ the bounds are a little less
stringent, i.e. $A_d \leq 1.3$ for $m_\smallh = 100$ GeV and $A_d \leq 4$
for $m_\smallh = 400$ GeV.
From the figure it is also apparent that the bound on $A_d$ for a
fixed value of $A_u$ is almost independent of the colored or colorless
nature of the charged Higgs, with the colored case showing only
slightly stronger bounds. However, as seen in the previous section,
the bounds on $A_u$ derived from $R_b$ are more dependent on the
nature of the Higgs, so that the allowed regions for the $A_d$
coefficient are in fact different for the color-singlet and
color-octet charged Higgs.
\begin{figure}[t]
\begin{center}
\epsfig{figure=BR_vs_Ad_full.eps,width=12cm}
\end{center}
\vspace{-1cm}
\caption{${\rm BR}(\Bsg)$ as a function of $A_d$ in the type-III (solid line)
and type-C (dashed line) models for $A_u=0.3$ and two different
values of $m_\smallh$. The horizontal dashed lines specify the $2 \sigma$
interval around the experimental value of ${\rm BR}(\Bsg)$.}
\label{fig6}
\end{figure}
The case of same sign between $A_u$ and $A_d$ has the peculiarity
that, as shown in fig.~\ref{fig6}, there are actually two ranges of
values of $A_d$ that fit inside the $2\sigma$ allowed band for
${\rm BR}(\Bsg)$. This is related to the fact that, in this case, the sign of
the charged-Higgs contribution to the Wilson coefficient of the
magnetic dipole operator, $C_7^H$, is opposite to the sign of the SM
contribution, $C_7^{\rm{\scriptscriptstyle SM}}$. Since ${\rm BR}(\Bsg)$ is
roughly proportional to $|C_7^H+C_7^{\rm{\scriptscriptstyle SM}}|^2$,
as $A_d$ increases the branching ratio goes to zero when $C_7^H\approx
-C_7^{\rm{\scriptscriptstyle SM}}$, and it goes back inside the
$2\sigma$-allowed band when $C_7^H\approx
-2\,C_7^{\rm{\scriptscriptstyle SM}}$. Thus, the two ranges of
possible values of $A_d$ differ by the sign of the amplitude ${\cal
A}(b \to s \gamma)$, basically the sign of the Wilson coefficient
$C_7$. Although $B\to X_{\!s}\, \gamma$ allows both ranges of values for $A_d$, there
are other observables that are sensitive to the sign of $C_7$, thus
selecting one of the two options. Among them, we cite ${\rm BR}(B \to
X_s \,l^+ l^-)$ \cite{GHM} and the isospin-breaking asymmetry that can
be constructed from the exclusive neutral and charged $ B \to K^*
\gamma$ decay modes \cite{KN,MS}. Although, for both observables,
neither the experimental result nor the theoretical prediction is at
the same level of accuracy as for ${\rm BR}(\Bsg)$, these observables still
give a compelling indication that the sign of ${\cal A}(b \to s
\gamma)$ is that of the SM contribution, thus eliminating the
large-$A_d$ option.
\section{Conclusions}
Minimal flavor violation is a very popular criterion that is used to
suppress FCNC effects in models with new particles at the TeV scale.
The enforcement of the MFV criterion to the simplest extension of the
SM, i.e.~a model with a second Higgs doublet, allows the possibility
of color-singlet or color-octet Higgs field. For both cases we have
considered the two-loop QCD corrections to the charged Higgs boson
contribution to $R_b$. We found that for all four types of 2HDM with
MFV, after the inclusion of the two-loop QCD corrections, the
prediction for this observable is practically insensitive to the the
choice of renormalization scheme for the top mass entering the
one-loop part of the calculation. Thus, the upper bound on the
coupling $|A_u|$ derived from $R_b$ is improved, which for the type-I
and type-II models translates in a more robust lower bound on $\tan
\beta$. We have also computed the ${\cal O}(\alpha_s)$ contributions to
the Wilson coefficients relevant to the process $B\to X_{\!s}\, \gamma$ for the
color-octet case. This was the last missing ingredient to obtain a
determination of ${\rm BR}(\Bsg)$ at the NLO level for all 2HDM with MFV.
After the inclusion of the NLO corrections it is found that, in the
region allowed by the present experimental results, the $B\to X_{\!s}\, \gamma$
transition is fairly insensitive to the colored or colorless nature of
the charged Higgs. Furthermore, in type-III and C models, the bounds
on the $A_d$ parameter that can be obtained from ${\rm BR}(\Bsg)$ and other
observables rule out the large-$A_d$ region, where effects
proportional to the bottom mass could become important.
\section*{Acknowledgments}
We thank H.~Haber for communications concerning ref.~\cite{HL}, and
P.~Gambino for useful discussions. One of us (G.D.) also thanks
M.~Nebot for his contribution in the early stage of this project.
This work was supported in part by an EU Marie-Curie Research Training
Network under contract MRTN-CT-2006-035505 (HEPTOOLS) and by ANR under
contract BLAN07-2\_194882.
\newpage
\begin{appendletterA}
\section*{Appendix: Analytical expressions for the NLO Wilson
coefficients}
In this appendix we report the analytic expressions for the functions
$G^H_{7,8}$ and $\Delta^H_{7,8}$ entering $\delta C^{(1)}_{7,8}$. We
find
\begin{eqnarray}
G_7^H(y) &= &
C_R^1\,C_F \left\{ A_d A_u^* y\left[ \frac{4(-3+7y-2y^2)}{3(y-1)^3}{\rm Li}_2
\left( 1 - \frac{1}{y} \right)+
\frac{8-14y-3y^2}{3(y-1)^4}\ln^2y \right. \right.
\nonumber \\ &&~~~~~~~~~~~~~~~~~~~~~
\left. +\frac{2(-3-y+12y^2-2y^3)}{3(y-1)^4}\ln y
+\frac{7-13y+2y^2}{(y-1)^3}\right]
\nonumber \\ &&~~~~~~~~~~~
+|A_u|^2 y\left[ \frac{y(18-37y+8y^2)}{6 (y-1)^4}{\rm Li}_2
\left( 1 - \frac{1}{y} \right)+
\frac{y(-14+23y+3y^2)}{6 (y-1)^5}\ln^2y \right.
\nonumber \\ &&~~~~~~~~~~~~~~~~~~~~~
+\frac{-50+251y-174y^2-192y^3+21y^4}{54 (y-1)^5}\ln y
\nonumber \\ &&~~~~~~~~~~~~~~~~~~~~ \left.
\left . +\frac{797-5436y+7569y^2-1202y^3}{648(y-1)^4}\right] \right\}
\nonumber \\
&+& C_R^1\,C_R^2 \left\{ A_d A_u^* y\left[ \frac{-19 +25 y}{18 (y-1)^2}{\rm Li}_2
\left( 1 - \frac{1}{y} \right)+
\frac{-25+33 y+15 y^2}{36 (y-1)^4}\ln^2y \right. \right.
\nonumber \\ &&~~~~~~~~~~~~~~~~~~~~~~
\left. +\frac{33-59 y+ 3 y^2}{18(y-1)^3}\ln y
+\frac{-8+31y}{36 (y-1)^2}\right]
\nonumber \\ &&~~~~~~~~~~~
+ |A_u|^2 y\left[ \frac{12-25y}{36 (y-1)^2}{\rm Li}_2
\left( 1 - \frac{1}{y} \right)+
\frac{-12+85y-108 y^2+3y^3}{72 (y-1)^5}\ln^2y \right.
\nonumber \\ &&~~~~~~~~~~~~~~~~~~~~~
~~~~-\frac{50+33y-195y^2+16y^3}{108 (y-1)^4}\ln y \left.
\left . +\frac{17-29y+4y^2}{18(y-1)^3}\right] \right\}~,
\label{higceq}
\end{eqnarray}
\begin{eqnarray}
\Delta_7^H(y) &= & C^1_R\,C_F \left\{ A_d A_u^* y\left[ \frac{21-47y+8y^2}{6
(y-1)^3}+\frac{-8+14y+3y^2}{3(y-1)^4}\ln y \right] \right.
\nonumber \\ && ~~~~~~~~~~~~ \left.
+|A_u|^2 y\left[ \frac{-31-18y+135y^2-14y^3}{
36(y-1)^4}+\frac{y(14-23y-3y^2)}{6(y-1)^5}\ln y \right] \right\} \nonumber \\
&+& C^1_R \,C_R^2\left\{ A_d A_u^* y\left[ \frac{1}{3 (y-1)}
-\frac{1}{3(y-1)^2}\ln y \right] \right.
\nonumber \\ && ~~~~~~~~~~~~ \left.
+|A_u|^2 y\left[- \frac{1+ y}{12(y-1)^2}+
\frac{y}{6(y-1)^3}\ln y \right] \right\}~,
\label{del7}
\end{eqnarray}
\begin{eqnarray}
G_8^H(y) &= &
C^1_R \left\{A_d A_u^* \frac13 y\left[ \frac{-36+25y-17y^2}{2(y-1)^3}{\rm Li}_2
\left( 1 - \frac{1}{y} \right)+
\frac{19+17y}{(y-1)^4}\ln^2y \right. \right.
\nonumber \\ &&~~~~~~~~~~~~~~~~~~~
\left. +\frac{-3-187y+12y^2-14y^3}{4(y-1)^4}\ln y
+\frac{3(143-44y+29y^2)}{8(y-1)^3}\right]
\nonumber \\ & &~~~~~~~
+|A_u|^2 \frac16 y\left[ \frac{y(30-17y+13y^2)}{(y-1)^4}{\rm Li}_2
\left( 1 - \frac{1}{y} \right)-
\frac{y(31+17y)}{(y-1)^5}\ln^2y \right.
\nonumber \\ &&~~~~~~~~~~~~~~~~~~~
+\frac{-226+817y+1353y^2+318y^3+42y^4}{36(y-1)^5}\ln y
\nonumber \\ && ~~~~~~~~~~~~~~~~~~~ \left.
\left. +\frac{1130-18153y+7650y^2-4451y^3}{216(y-1)^4}\right] \right\}
\nonumber \\
&+& C^2_R \left\{A_d A_u^* \frac19 y\left[ \frac{-43+34y}{4(y-1)^2}{\rm Li}_2
\left( 1 - \frac{1}{y} \right)+
\frac{-157-108y +81 y^2}{8(y-1)^4}\ln^2y \right. \right.
\nonumber \\ &&~~~~~~~~~~~~~~~~~~~
\left. +\frac{-51-22y+57y^2}{8(y-1)^3}\ln y
+\frac{5(13-8y)}{(y-1)^2}\right]
\nonumber \\ & &~~~~~
+|A_u|^2 \frac1{144} y\left[ \frac{-15+149y-122y^2}{(y-1)^3}{\rm Li}_2
\left( 1 - \frac{1}{y} \right) \right.
\nonumber \\ &&~~~~~~~~~~~~~~~~~~~~
-\frac{15-533y - 237 y^2 + 243 y^3}{2 (y-1)^5}\ln^2y
\nonumber \\
&&~~~~~~~~~~~~~~~~~~~~
-\frac{172-744y+357y^2+23y^3}{3(y-1)^4}\ln y
\left. \left. -\frac{203+1174y-737y^2}{2(y-1)^3}\right] \right\}
\end{eqnarray}
\begin{eqnarray}
\Delta_8^H(y) &= & C^1_R \left\{ A_d A_u^* \frac{1}{3}y
\left[ \frac{81-16y+7y^2}{2(y-1)^3}-\frac{19+17y}{(y-1)^4}\ln y \right] \right.
\nonumber \\ &&~~~~~~~
\left. +|A_u|^2\frac{1}{6}y\left[ \frac{-38-261y+18y^2-7y^3}{
6(y-1)^4}+\frac{y(31+17y)}{(y-1)^5}\ln y \right] \right\}
\nonumber \\
&+& C^2_R \left\{ A_d A_u^* \frac{1}{6}y\left[ \frac{31-7y}{(y-1)^2}
-\frac{19+5y}{(y-1)^3}\ln y \right] \right.
\nonumber \\ &&~~~~~~
\left. +|A_u|^2\frac{1}{12}y\left[ \frac{-19-60y+7y^2}{2(y-1)^3}
+\frac{y(31+5y)}{(y-1)^4}\ln y \right] \right\}
\label{del8}
\end{eqnarray}
\begin{eqnarray}
E^H(y)& = & C^1_R \, |A_u|^2 \frac16 y
\left[ \frac{16-29y+7y^2}{6(y-1)^3}+\frac{-2+3y}{(y-1)^4} \ln y \right]
\nonumber \\
&+& C^2_R \, |A_u|^2 \frac14 y
\left[ \frac{-1}{(y-1)}+\frac{2+y}{3 (y-1)^2} \ln y \right]~.
\end{eqnarray}
The results for type I-III models are recovered setting $C^1_R =1$ and
$C^2_R =0$, while the result for the case of a colored charged scalar
in the adjoint of $SU(3)$ is obtained with $C^1_R =C_F$ and $C^2_R
=N_c$.
\vspace*{0.5cm}
\paragraph{\Large Note added:}
after the publication of our paper, we became aware of
ref.~\cite{bobeth}, which contains two-loop formulae for the Wilson
coefficients of the operators relevant to $b\rightarrow s \gamma$ in a
generic model with a heavy fermion and a heavy scalar. While
ref.~\cite{bobeth} does not specifically discuss the case of the
type-C 2HDM, the results presented in our appendix can be reproduced
with appropriate substitutions in the results of ref.~\cite{bobeth},
taking into account the different renormalization conditions on the
charged-Higgs mass. We thank Christoph Bobeth for performing this
useful comparison.
\end{appendletterA}
\newpage
|
\section*{Supplementary Material}
\subsection{Sufficient Conditions' Comparison**}
We briefly compare the results for reg-mod-BPDN, mod-BPDN and BPDN, primarily by comparing the sufficient conditions required for them to hold. The comparison of the bounds is not easy since each holds under a different set of sufficient conditions. This will be done later using the results of Section IV which hold without any sufficient conditions. For the comparison of sufficient conditions, we use the restricted isometry constant (RIC), $\delta_S$ and restricted orthogonality constant (ROC), $\theta_{S,S'}$ \cite{decodinglp} defined next. These depend only on the sizes of the sets $T$, $\Delta$ and $N$ and hence make a theoretical comparison easier. However the comparison can only be qualitative. The RIC and ROC are not computable (computation complexity is exponential in the set size) and hence cannot be used for numerical comparisons.
On the other hand, the ERC and the bounds obtained based on the ERC approach are computable and can be used for a quantitative numerical comparison. We do this comparison later in Sec. V.
\subsubsection{RIC and ROC definition}
The $S$ restricted isometry constant (RIC) \cite{decodinglp}, $\delta_S$, for a matrix, $A$, is defined as the
smallest positive number satisfying
\begin{equation}
(1- \delta_S) \|c\|_2^2 \le \|A_T c\|_2^2 \le (1 + \delta_S) \|c\|_2^2
\label{def_delta}
\end{equation}
for all subsets $T$ of cardinality $|T| \le S$ and all real vectors $c$ of length $|T|$.
The restricted orthogonality constant (ROC) \cite{decodinglp}, $\theta_{S_1,S_2}$, is defined as the smallest real number satisfying
\begin{equation}
| {c_1}'{A_{T_1}}'A_{T_2} c_2 | \le \theta_{S_1,S_2} \|c_1\|_2 \|c_2\|_2
\label{def_theta}
\end{equation}
for all disjoint sets $T_1, T_2 $ with $|T_1| \le S_1$, $|T_2| \le S_2$ and $S_1+S_2 \le m$, and for all nonzero vectors $c_1$, $c_2$ of length $|T_1|$, $|T_2|$ respectively.
If we let $c_2 = {A_{T_2}}'A_{T_1} c_1$, then (\ref{def_theta}) implies that $\|{A_{T_2}}'A_{T_1} c_1\|_2^2 \le \theta_{S_1,S_2} \|c_1\|_2 \|{A_{T_2}}'A_{T_1} c_1\|_2$ and so $\|{A_{T_2}}'A_{T_1} c_1\|_2 \le \theta_{S_1,S_2} \|c_1\|_2$ for all nonzero $c_1$'s.
Thus, using the definition of the matrix norm $\|M\|_2$, this implies that
\begin{equation}
\|{A_{T_2}}'A_{T_1}\|_2 \le \theta_{S_1,S_2}
\label{theta_bnd}
\end{equation}
for all sets with $|T_1| \le S_1$, $|T_2| \le S_2$.
\subsubsection{Sufficient Conditions' Comparison}
Consider mod-BPDN versus BPDN first. Let us compare their ERC's.
Using
(\ref{theta_bnd}), $\|({A_T}'A_T + \lambda I_T)^{-1}\|_2 \le
1/(1-\delta_{|T|} + \lambda)$ and the fact that for a vector $z$ of length $l$, $\|z\|_1 \le \sqrt{l} \|z\|_2$,%
\begin{eqnarray}
ERC_{T,\lambda}(\Delta) & \ge & 1-\sqrt{|\Delta|}\|P_{T,\lambda}(\Delta)\|_2 \|{A_{\Delta}}'M_{T,\lambda}A_{\omega}\|_2 \nonumber \\
& \ge & 1- {\sqrt{|\Delta|}} \frac{(\theta_{|\Delta|,1}+\frac{\theta_{|\Delta|,|T|}\theta_{|\Delta|,1}}{1-\delta_{|T|}+\lambda})} {1- \delta_{|\Delta|} - \frac{\theta_{|T|,|\Delta|}^2}{1-\delta_{|T|}+\lambda} }
\end{eqnarray}
where the numerator of the second term comes from bounding $\|{A_{\Delta}}'M_{T,\lambda}A_{\omega}\|_2$ and the
denominator of the second term comes from bounding $\|P_{T,\lambda}(\Delta)\|_2$.
In practice, for example in recursive reconstruction applications like real-time dynamic MRI,
usually $|\Delta| \approx |\Delta_e| \ll |N|$ and $|N| \approx |T| \approx |T \cup \Delta|$ \cite{modcsMRI}.
Under this assumption, when fewer measurements are available (but still enough to ensure that $\delta_{|N|}<1$), the denominator for the second term of $ERC_{\emptyset,0}(N)$ (BPDN),
$1-\delta_{|N|}$, will be smaller than that of $ ERC_{T,0}(\Delta)$ (mod-BPDN),
$1-\delta_{|\Delta|} - \frac{\theta_{|T|,|\Delta|}^2}{1-\delta_{|T|}}$. Also, $\sqrt{|N|}$ in its numerator
will be larger than $\sqrt{|\Delta|}$ for mod-BPDN, while the other numerator terms will be similar in both cases. This can result in a smaller (and possibly negative)
lower bound on the ERC for BPDN.
To compare reg-mod-BPDN and mod-BPDN, notice that mod-BPDN needs $A_{T\cup \Delta}$ to be full rank where as reg-mod-BPDN only needs $A_{\Delta}$ to be full rank which is much weaker.
We show a numerical comparison in Table I (simulation details given in Sec. V). Notice that BPDN needs $90\%$ measurements for its ERC to become positive where as mod-BPDN only needs $19\%$. Moreover even with $90\%$ measurements, its ERC is just positive and very small. As a result its error bound is large ($27\%$ normalized mean squared error (NMSE)).
Similarly notice that mod-BPDN needs $n \ge 19\%$ while for reg-mod-BPDN $n=13\%$ also suffices
\begin{remark}
A sufficient conditions' comparison only provides a comparison of when a given result can be applied to provide a bound on the reconstruction error.
For example, in simulations, of course BPDN provides a good reconstruction using much lesser than 90\% measurements. However, when $n<90\%$ we cannot
bound its reconstruction error using Theorem 1 above (for BPDN this is the same as the result of \cite{justrelax}).
\end{remark}
\subsection{Equivalence between Theorem 2 and Theorem 3 bounds}\label{AppLLN}
We can use the weak law of large numbers (WLLN) to argue that as $n$, $s\triangleq |N|$ approach to infinity the bound from Theorem 3 converges to that of Theorem 2 in probability. We give the basic idea here. The complete proof will be in future work. The WLLN argument applies when
\begin{itemize}
\item Each element of $A$ is iid with zero mean and variance $1/n$, i.e. $A = \frac{1}{\sqrt{n}}Z$ where each element of $Z$ is iid with zero mean and unit variance.
\item The noise $w$ is bounded in $\ell_2$ norm, i.e. $\|w\|_2 \le \eta$ and
\item $n, s \rightarrow \infty$
\end{itemize}
WLLN can be used to argue that as $n,s \rightarrow \infty$, with high probability (w.h.p.), $ERC(\tDelta)$ and the multipliers $g_1, g_2, g_3,g_4$ depend only on the size, $k$, of the set $\tDelta$, i.e. they are the same for all sets $\tDelta$ of a given size. Thus, the only term in $g(\tDelta)$ that varies for different sets $\tDelta \in \mathcal{G}_k$ is $\|x_{\Delta \setminus \tDelta}\|_2$. Thus $\arg\min_{\mathcal{G}_k} g(\tDelta) = \arg\min_{\mathcal{G}_k} \|x_{\Delta \setminus \tDelta}\|_2$.
Since $ERC$ also only depends on $k$, for a given $k$, either $ERC(k) > 0$ or $ERC(k) < 0$. When $ERC(k) > 0$, $\mathcal{G}_k=\{\tDelta \subseteq \Delta, |\tDelta|=k\}$, where as when $ERC(k) < 0$, $\mathcal{G}_k$ is empty. The minimum value over an empty set is infinity. Thus, $\min_{\mathcal{G}_k} \|x_{\Delta \setminus \tDelta}\|_2 =B_k$. Using (\ref{regmodBPDNmediumtightbound}), (\ref{Gkdef}) and (\ref{kmindef}), this means that $g(\tDelta^*)=g(\tDelta^{**})$, i.e. the bounds from Theorems 2 and 3 are equal.
The WLLN argument is as follows.
Note that all terms in $g_1,g_2,g_3,g_4$ and $ERC$ that depend on $\tDelta$ are functions of either ${A_\tDelta}'A_\tDelta$ or ${A_T}'A_\tDelta$ or ${A_\tDelta}'M_{T,\lambda} A_\tDelta$ or $A_{T\cup \tDelta}'w$
Consider ${A_\tDelta}'A_\tDelta$.
\begin{eqnarray}
\hspace{-1mm}({A_\tDelta}'A_\tDelta)_{i,j} = \left\{ \begin{array}{ll}
\sum_{r=1}^n A_{i, r}^2 = \frac{1}{n} \sum_{r=1}^n Z_{i, r}^2 & \ \text{if} \ i=j \\
\sum_{r=1}^n A_{i, r} A_{j,r} = \frac{1}{n} \sum_{r=1}^n Z_{i, r} Z_{j,r} & \ \text{if} \ i \neq j
\end{array}
\right. \nonumber
\end{eqnarray}
Clearly $\mathbb{E}[Z_{i, r}^2] = 1$ and its variance, $Var[Z_{i, r}^2] = 3$ where as $\mathbb{E}[Z_{i, r} Z_{j,r}] = 0$
while $Var[Z_{i, r} Z_{j,r}] = 1$. Here, $\mathbb{E}[\cdot]$ and $Var[\cdot]$ denote the expectation and variance computed over the distribution of $A$. Thus by WLLN, as $n \rightarrow \infty$, ${A_\tDelta}'A_\tDelta$ approaches the
identity matrix, $I_k$ w.h.p.. A similar argument can be made for each element of ${A_T}'A_\tDelta$ to show that this
approaches the zero matrix as $n \rightarrow \infty$. A similar argument can also be made for $M_{T,\lambda}$
when $s:=|N|$ (and hence $|T|$) goes to infinity to show that all its diagonal elements converge to one value and
all the non-diagonal ones converge to another value. This fact can then be used to make a WLLN argument
for each element of ${A_{\tDelta}}'M_{T,\lambda} A_{\tDelta}$. Now consider $g_4$ which contains the term $\|{A_{(T\cup \Delta)^c} }' w\|_{\infty}$. Notice that
$({A_{(T\cup \Delta)^c }}' w)_i = \sum_{j=1}^n w_j A_{j,i}$. Taking expectations only over the elements of $A$, $\mathbb{E}[({A_{(T\cup \Delta)^c }}' w)_i] = 0$ and $Var[({A_{(T\cup \Delta)^c }}' w)_i ] = \sum_{j=1}^n w_j^2 \frac{1}{n} \le \frac{\eta^2}{n}$. Thus by WLLN, each element of the vector ${A_{(T\cup \Delta)^c }}' w$ approaches zero, and hence its infinity norm also approaches zero w.h.p..
Thus, w.h.p., for a given size $k$, all these three matrices and $\|{A_{(T\cup \Delta)^c}}'w\|_{\infty}$, and as a result
all of $ERC,g_1,g_2,g_3, g_4$, converge to a value that does not depend on the set $\tDelta$.
\section{Numerical Experiments}
In this section, we show both upper bound comparisons and actual reconstruction error comparisons.
The upper bound comparison only tells us that the performance guarantees of reg-mod-BPDN are better than those for the other methods. To actually demonstrate that reg-mod-BPDN outperforms the others, we need to compare the actual reconstruction errors.
Based on useful comments by anonymous reviewers, we have reorganized this section. After giving the simulation model, we first show the reconstruction error comparisons for recovering simulated sparse signals from random Gaussian measurements. Next, we show comparisons for recursive dynamic MRI reconstruction of a larynx image sequence. In this comparison, we also show the usefulness of the Theorem 3 in helping us select a good value of $\gamma$. In the last three subsections, we show numerical comparisons of the results of the various theorems. The upper bound comparisons of Theorem 3 and the comparison of the corresponding reconstruction errors suggests that the bounds for reg-mod-BPDN and BPDN are tight under the scenarios evaluated. Hence, they can be used as a proxy to decide which algorithm to use when. We show this for both random Gaussian and partial Fourier measurements.
\subsection{Simulation Model}
The notation $z = \pm a$ means that we generate each element of the vector $z$ independently and each is either $+a$ or $-a$ with probability 1/2. The notation $\nu \sim \n(0, \Sigma)$ means that $\nu$ is generated from a Gaussian distribution with mean 0 and covariance matrix $\Sigma$. We use $\lfloor a \rfloor$ to denote the largest integer less than or equal to $a$. Independent and identically distributed is abbreviated as iid. Also, N-RMSE refers to the normalized root mean squared error.
We use the recursive reconstruction application \cite{LSCS,modcsjournal} to motivate the simulation model. In this case,
assuming that slow support and slow signal value change hold [see Fig. \ref{slowchange}], we can use the reconstructed value of the signal
at the previous time as $\muhat$ and its support as $T$. To simulate the effect of slow signal value change, we let $x_N = \mu_N + \nu$ where $\nu$
is a small iid Gaussian deviation and we let $\muhat_{T \cap N} = \mu_{T \cap N}$ (and so $x_{T \cap N} = \muhat_{T \cap N} + \nu_{T \cap N}$)
The extras' set, $\Delta_e = T \setminus N$, contains elements that got removed from the support at the current time or at a few previous times (but so far did not get removed from the support estimate). In most practical applications, only small valued elements at the previous time get removed from the support and hence the magnitude of $\hat\mu$ on $\Delta_e$ will be small. We use $\beta_s$ to denote this small magnitude, i.e. we simulate $(\muhat)_{\Delta_e} = \pm \beta_s$.%
The misses' set at time $t$, $\Delta$, definitely includes the elements that just got added to the support at $t$ or the ones that previously got added but did not get detected into the support estimate so far. The new elements typically get added at a small value and their value slowly increases to a large one. Thus, elements in $\Delta$ will either have small magnitude (corresponding to the current newly added ones), or will have larger magnitude but still smaller than that of elements already in $N \cap T$.
To simulate this, we do the following. (a) We simulate the elements on $N \cap T$ to have large magnitude, $\beta_l$, i.e.
we let $(\mu)_{N \cap T} = \pm \beta_l$.
(b) We split the set $\Delta$ into two disjoint parts, $\Delta_1$ and $\Delta_2 = \Delta \setminus \Delta_1$. The set $\Delta_1$ contains the small (e.g. newly added) elements, i.e. $(\mu)_{\Delta_1} = \pm \beta_s$.
The set $\Delta_2$ contains the larger elements, though still with magnitudes smaller than those in $N \cap T$, i.e. $(\mu)_{\Delta_2} = \pm \beta_m$, where $\beta_l \ge \beta_m \ge \beta_s$
In summary, we use the following simulation model.
\begin{eqnarray}
(x)_N &=& (\mu)_N + \nu, \ \ \nu \sim \n(0,\sigma_p^2 I) \nonumber \\
(x)_{N^c} &=& 0 \label{xgen} \\
\text{where} \ \ (\mu)_{N \cap T} &=& \pm \beta_l \nonumber \\
(\mu)_{\Delta_1} &=& \pm \beta_s, \ \ (\mu)_{\Delta_2} = \pm \beta_m \nonumber \\
(\mu)_{N^c} &=& 0 \label{mugen}
\end{eqnarray}
and
\begin{eqnarray}
(\hat\mu)_{T \cap N} &=& (\mu)_{T \cap N} = \pm \beta_l \nonumber \\
(\hat\mu)_{\Delta_e} &=& \pm \beta_s \nonumber \\
(\hat\mu)_{T^c} &=& 0 \label{muhatgen}
\end{eqnarray}
We generate the support of $x$, $N$, of size $|N|$, uniformly at random from $[1,...,m]$. We generate $\Delta$ with size $|\Delta|$ and $\Delta_e$ with size $|\Delta_e|$ uniformly at random from $N$ and from $N^c$ respectively. The set $\Delta_1$ of size $|\Delta_1|=\lfloor |\Delta|/2\rfloor$ is generated uniformly at random from $\Delta$. The set $\Delta_2 = \Delta \setminus \Delta_1$. We let $T = N \cup \Delta_e \setminus \Delta$. We generate $\mu$ and then $x$ using (\ref{mugen}) and (\ref{xgen}). We generate $\hat\mu$ using (\ref{muhatgen}).
In some simulations, we simulated the more difficult case where $\beta_m = \beta_s$. In this case, all elements on $\Delta$ were identically generated and hence we did not need $\Delta_1$
\begin{figure*}[t!]
\subfigure[\small{$n=0.13m$,$\sigma_p^2=10^{-3}$,$\sigma_w^2=10^{-5}$}]{
\includegraphics [width=9cm,height=6cm]{regmodBPDNreconmeas13smallnoisegoodprioralllatestmodel.eps}
}
\hspace{-5mm}
\subfigure[\small{$n=0.13m$,$\sigma_p^2=10^{-1}$,$\sigma_w^2=10^{-5}$}]{
\includegraphics [width=9cm,height=6cm]{regmodBPDNreconmeas13smallnoisebadprioralllatestmodelnokfcs.eps}
}\\
\subfigure[\small{$n=0.3m$,$\sigma_p^2=10^{-3}$,$\sigma_w^2=10^{-4}$}]{
\includegraphics [width=9cm,height=6cm]{regmodBPDNreconmeas30mediumnoisegoodprioralloldmodelnokfcs.eps}
}
\hspace{-5mm}
\subfigure[\small{$n=0.3m$,$\sigma_p^2=10^{-1}$,$\sigma_w^2=10^{-5}$}]{
\includegraphics [width=9cm,height=6cm]{regmodBPDNreconmeas30smallnoisebadprioralloldmodelnokfcs.eps}
}
\vspace{-1mm}
\caption{\small{The N-RMSE for reg-mod-BPDN, mod-BPDN, BPDN, LS-CS, KF-CS, weighted $\ell_1$, CS-residual, CS-mod-residual and modified-CS-residual are plotted.
For $n=0.13m$ , reg-mod-BPDN has smaller errors than those of mod-BPDN and the gap is larger when the signal estimate is good. For $n=0.3m$, the errors of reg-mod-BPDN, mod-BPDN and weighted $\ell_1$ are close and all small. }
} \label{reconfig}
\end{figure*}
\subsection{Reconstruction Error Comparisons}
In Fig. \ref{reconfig}, we compare the Monte Carlo average of the reconstruction error of reg-mod-BPDN with that of mod-BPDN, BPDN, weighted $\ell_1$ \cite{weightedl1} given in (\ref{weightedl1}),
CS-residual given in (\ref{CSresidual}), CS-mod-residual given in (\ref{CSmodresidual}) and modified-CS-residual\cite{modcsMRI} given in (\ref{modCSresidual}). Simulation was done according to the model specified above. We used random Gaussian measurements in this simulation, i.e. we generated $A$ as an $n \times m$ matrix with iid zero mean Gaussian entries and normalized each column to unit $\ell_2$ norm.
We experimented with two choices of $n$, $n=0.13m$ (where reg-mod-BPDN outperforms mod-BPDN) and $n=0.3m$ (where both are similar) and two values of $\sigma_p^2$, $\sigma_p^2=0.001$ (good prior) and $\sigma_p^2=0.1$ (bad prior). For the cases of Fig \ref{reconfig}(a) ($n=0.13m$, $\sigma_p^2=0.001$) and Fig \ref{reconfig}(b) ($n=0.13m$, $\sigma_p^2=0.1$), we used signal length $m=256$, support size $|N|=0.1m=26$ and support extras' size, $|\Delta_e|=0.1|N|$. The misses' size, $|\Delta|$, was varied between 0 and $0.2|N|$ (these numbers were motivated by the medical imaging application, we used larger numbers than what are shown in Fig. \ref{slowchange}). We used $\beta_l=1$, $\beta_m=0.4$ and $\beta_s=0.2$. The noise variance was $\sigma_w^2=10^{-5}$.
For the last two figures, Fig \ref{reconfig}(c) ($n=0.3m$, $\sigma_p^2=0.001$) and Fig \ref{reconfig}(d) ($n=0.3m$, $\sigma_p^2=0.1$), for which $n$ was larger, we used $\beta_m= \beta_s = 0.25$ which is a more difficult case for reg-mod-BPDN. For Fig. \ref{reconfig}(c), we also used a larger noise variance $\sigma_w^2=10^{-4}$. All other parameters were the same.
{For applications where some training data is available, $\gamma$ and $\lambda$ for reg-mod-BPDN can be chosen by interpreting the reg-mod-BPDN solution as the maximum a posteriori (MAP) estimate under a certain prior signal model} (assume $x_T$ is Gaussian with mean $\hat{\mu}_T$ and variance $\sigma_p^2$ and $x_{T^c}$ is independent of $x_T$ and is iid Laplacian with parameter $b$). This idea is explained in detail in \cite{modcsjournal}. However, there is no easy way to do this for the other methods. Alternatively, choosing $\gamma$ and $\lambda$ according to Theorem 3 gives another good start point. We can do this for mod-BPDN and BPDN, but we cannot do this for the other methods (we show examples using this approach later).
For a fair error comparison, for each algorithm, we selected $\gamma$ from a set of values $[0.00001\ 0.00005\ 0.0001\ 0.0005\ 0.001\ 0.005\ 0.01\ 0.1]$. We tried all these values for a small number of simulations (10 simulations) and then picked the best one (one with the smallest N-RMSE) for each algorithm. For weighted $\ell_1$ reconstruction, we also pick the best $\gamma'$ in (\ref{weightedl1}) from the same set in the same way\footnote{To give an example, our finally selected numbers for Fig. \ref{reconfig}(d) were $\gamma=0.01,0.001, 0.001,0.001,0.001,0.001,0.01,0.01$ for BPDN, mod-BPDN, reg-mod-BPDN, weighted $\ell_1$, LS-CS, CS-residual, CS-mod-residual, mod-CS-residual respectively and $\gamma'=0.0001$}. For reg-mod-BPDN, $\lambda$ should be larger when the signal estimate is good and should be decreased when the signal estimate is not so good. We can use $\lambda=\alpha \sigma_w^2 / \sigma_p^2$ to adaptively determine its value for different choices of $\sigma_w^2$ and $\sigma_p^2$. In our simulations, we used $\alpha = 0.2$ for Fig. \ref{reconfig} (a), (b) and (d) and $\alpha = 0.05$ for Fig. \ref{reconfig}(c).
We fixed the chosen $\gamma$, $\gamma'$ and $\lambda$ and did Monte Carlo averaging over 100 simulations. We conclude the following. (1) When the signal estimate is not good (Fig. \ref{reconfig}(b),(d)) or when $n$ is small (Fig. \ref{reconfig}(a),(b)), CS-residual and CS-mod-residual have significantly larger error than reg-mod-BPDN. (2) In case of Fig. \ref{reconfig}(d) ($n=0.3m$), they also have larger error than mod-BPDN.
(3) In all four cases, weighed $\ell_1$ and mod-BPDN have similar performance. This is also similar to that of reg-mod-BPDN in case of $n=0.3m$, but is much worse in case of $n=0.13m$.
(4) We also show a comparison with regmodBPDN-var in Fig. \ref{reconfig}(a). Notice that it has larger errors than reg-mod-BPDN for reasons explained in Sec. I-C.
\begin{figure
\center
\includegraphics [width=9cm,height=6cm]{larynxreconusingThm3boundregcsrescsregmapcsmapcsresmap.eps}
\vspace{-4mm}
\caption{\small{Reconstructing a $32 \times 32$ block of the actual (compressible) larynx sequence from partial Fourier measurements.
Measurements $n=0.18m$ for $t=0$ and $n=0.06m$ for $t>0$.
Reg-mod-BPDN has the smallest reconstruction error among all methods.}
} \label{larynxcompareregmodBPDN}
\end{figure}
\subsection{Dynamic MRI application using $\gamma$ from Theorem 3}
In Fig. \ref{larynxcompareregmodBPDN}, we show comparisons for simulated dynamic MR imaging of an actual larynx image
sequence (Fig. \ref{slowchange} (a)(i)). The larynx image is not exactly sparse but is only compressible in the wavelet domain.
We used a two-level Daubechies-4 2D discrete wavelet transform (DWT). The $99$\%-energy support size of its wavelet transform vector,
$|N_t|\approx 0.07m$. Also, $|\Delta_t|\approx 0.001m$ and $|\Delta_{e,t}|\approx 0.002m$. We used a $32\times 32$ block of this sequence and at
each time and simulated undersampled MRI, i.e. we selected $n$ 2D discrete Fourier transform (DFT) coefficients using the variable density
sampling scheme of \cite{sparseMRI}, and added iid Gaussian noise with zero mean and variance $\sigma_w^2=10$ to each of them. Using a small $32\times 32$ block allows easy implementation using CVX (for full sized image sequences, one needs specialized code). We used $n_0=0.18m$ at $t=0$ and $n=0.06m$ at $t>0$.
We implemented dynamic reg-mod-BPDN as described in Algorithm 1. In this problem, the matrix $A=F_{u}\cdot W^{-1}$ where $F_u$ contains the selected rows of the 2D-DFT matrix and $W$ is the inverse 2D-DWT matrix (for a two-level Daubechies-4 wavelet). Reg-mod-BPDN was compared with similarly implemented reg-mod-BPDN-var and CS-residual algorithms (CS-residual only solved simple BPDN at $t=0$). We also compared with simple BPDN (BPDN done for each frame separately). For reg-mod-BPDN and reg-mod-BPDN-var, the support estimation threshold, $\rho$, was chosen as suggested in \cite{modcsjournal}: we used $\rho=20$ which is slightly larger than the smallest magnitude element in the $99\%$-energy support which is $15$. At $t=0$, we used $T_0$ to be the set of indices of the wavelet approximation coefficients.
To choose $\gamma$ and $\lambda$ we tried two different things. (a) We used $\lambda$ and $\gamma$ from the set
$[0.00001\ 0.00005\ 0.0001\ 0.0005\ 0.001\ 0.005\ 0.01\ 0.1]$ to do the reconstruction for a short training sequence (5 frames),
and used the average error to pick the best $\lambda$ and $\gamma$. We call the resulting reconstruction error plot reg-mod-BPDN-opt.
(b) We computed the average of the $\gamma^*$ obtained from Theorem 3 for the 5-frame training sequence and used this as $\gamma$ for the test sequence.
We selected $\lambda$ from the above set by choosing the one that minimizes the average of the bound of Theorem 3 for the 5 frames.
We call the resulting error plot reg-mod-BPDN-$\gamma^*$. The same two things were also done for BPDN and CS-residual as well. For reg-mod-BPDN-var,
we only did (a).
From Fig. \ref{larynxcompareregmodBPDN}, we can conclude the following. (1) Reg-mod-BPDN significantly outperforms the other methods when using so few measurements. (2) Reg-mod-BPDN-var and reg-mod-BPDN have similar performance in this case. (3) The reconstruction performance of reg-mod-BPDN using $\gamma^*$ from Theorem 3 is close to that of reg-mod-BPDN using the best $\gamma$ chosen from a large set. This indicates that Theorem 3 provides a good way to select $\gamma$ in practice.
\subsection{Comparing the result of Theorem 1}
In Table I, we compare the result of Theorem 1 for reg-mod-BPDN, mod-BPDN and BPDN.
We used $m=256$, $|N|=26=0.1m$, $|\Delta|=0.04|N| = |\Delta_e|$, $\sigma_p^2=10^{-3}$, $\beta_l=1$ and $\beta_m=\beta_s=0.25$. Also, $\sigma_w^2=10^{-5}$ and we varied $n$. For each experiment with a given $n$, we did the following. We did $100$ Monte Carlo simulations. Each time, we evaluated the sufficient conditions for the bound of reg-mod-BPDN to hold. We say the bound {\em holds} if all the sufficient conditions hold for at least $98$ realizations. If this did not happen, we record {\em not hold} in Table I. If this did happen, then we recorded $\sqrt{\frac{\mathbb{E}[\text{bound}^2]}{\mathbb{E}[\|x\|^2_2]}}$ where $\mathbb{E}[\cdot ]$ denotes the Monte Carlo average computed over those realizations for which the sufficient conditions do hold. Here, ``bound" refers to the right hand side of (\ref{loosebound}) computed with $\gamma = \gamma^*_{T,\lambda}(\Delta)$ given in (\ref{regmodBPDNcond}). An analogous procedure was followed for both mod-BPDN and BPDN.
The comparisons are summarized in Table I. For reg-mod-BPDN, we selected $\lambda$ from the set\\
$[0.00001 \ 0.00005 \ 0.0001 \ 0.0005 \ 0.001 \ 0.005 \ 0.01 \ 0.1]$ by picking the one that gave the smallest bound. Clearly the reg-mod-BPDN result holds with the smallest $n$, while the BPDN result needs a very large $n$ ($n \ge 90\%$). Also even with $n=90\%$, the BPDN error bound is very large.
\begin{table}[h]
\center
\begin{tabular}{|c|c|c|c|}
\hline
$n$ & Reg-mod-BPDN & Mod-BPDN & BPDN \\
\hline
$0.13m$ & $0.885$ & not hold & not hold \\
\hline
$0.19m$ & $0.161$ & $0.303$ & not hold \\
\hline
$0.5m$ & $0.0199$ & $0.0199$ & not hold \\
\hline
$0.9m$ & $0.014$ & $0.014$ & $0.27$ \\
\hline
\end{tabular}\label{tableb}
\caption{\small{Sufficient conditions and normalized bounds comparison of reg-mod-BPDN, mod-BPDN and
BPDN. Signal length $m=256$, support size $|N|=0.1m$, $|\Delta|=4\%|N|$, $\Delta_e=4\%|N|$, $\sigma_w^2=10^{-5}$ and $\sigma_p^2=10^{-3}$. ``not hold" means the one or all of the sufficient conditions does not hold.}}
\end{table}
\subsection{Comparing Theorems 1, 2, 3}
In Fig. \ref{bound4comparefig} (a), we compare the results from Theorems 1, 2 and 3 for one simulation. We plot $\frac{\text{bound}}{\|x\|_2}$ for $|\Delta|/|N|$ ranging from 0 to 0.2. Also, we used $m=256$, $|N|=26$, $|\Delta_e|=0.1|N|$, $\sigma_p^2=10^{-3}$, $\beta_l=1$ and $\beta_m=\beta_s=0.25$. Also, $n=0.13m$ and $\sigma_w^2=10^{-5}$. We used $\gamma = \gamma^*$ given in the respective theorems, and we set $\lambda= 10 \sigma_w^2 / \sigma_p^2$.
We notice the following. (1) The bound of Theorem 1 is much larger for than that of Theorem 2 or 3, even for $|\Delta|=0.04|N|$. (2) For larger values of $|\Delta|$, the sufficient conditions of Theorem 1 do not hold and hence it does not provide a bound at all. (3) For reasons explained in Sec. IV, in this case, the bound of Theorem 3 is equal to that of Theorem 2. Recall that the computational complexity of the bound from Theorem 2 is exponential in $|\Delta|$. However if $|\Delta|$ is small, e.g. in our simulations $|\Delta| \le 5$, this is still doable.
\begin{figure*}[t!]
\centerline{
\hspace{-8mm}
\subfigure[$n=0.13m$,$\sigma_p^2=10^{-3}$,$\sigma_w^2=10^{-5}$]{
\includegraphics [width=6.1cm,height=4.8cm]{regmodBPDN13meassmallnoisegoodprior3newbounds.eps}
}
\hspace{-8mm}
\subfigure[$n=0.13m$,$\sigma_p^2=10^{-3}$,$\sigma_w^2=10^{-5}$]{
\includegraphics [width=6.1cm,height=4.8cm]{regmodBPDNboundmeas13smallnoisegoodprionewrtightboundandreconmc100oldmodel.eps}
}
\hspace{-8mm}
\subfigure[$n=0.17m$,$\sigma_p^2=10^{-3}$,$\sigma_w^2=10^{-3}$]{
\includegraphics [width=6.1cm,height=4.8cm]{regmodBPDNboundreconThm3partialFTmeas17goodprior.eps}
}
}
\vspace{-1mm}
\caption{\small{In (a), we compare the three bounds from Theorem 1, 2 and 3 for one realization of $x$.
In (b) and (c), we compare the normalized average bounds from Theorem 3 and reconstruction errors with random Gaussian and
partial Fourier measurements respectively.
} \label{bound4comparefig}
\end{figure*}
\subsection{Upper bound comparisons using Theorem 3}
In Fig. \ref{bound4comparefig}(b), we do two things. (1) We compare the reconstruction error bounds from Theorem 3 for reg-mod-BPDN, mod-BPDN and BPDN and compare them with the bounds for LS-CS error given in \cite[Corollary 1]{LSCSbound}. All bounds hold without any sufficient conditions which is what makes this comparison possible.
(2) We also use the $\gamma^*$ given by Theorem 3 to obtain the reconstructions and compute the Monte Carlo averaged N-RMSE. Comparing this with the Monte Carlo averaged upper bound on the N-RMSE,$\sqrt{\frac{\mathbb{E}[\text{bound}^2]}{\mathbb{E}[\|x\|^2_2]}}$, allows us to evaluate the tightness of a bound. Here $\mathbb{E}[\cdot]$ denotes the mean computed over 100 Monte Carlo simulations and ``bound" refers to the right hand side of (\ref{regmodBPDNmediumtightbound}).
We used $m=256$, $|N|=26$, $|\Delta_e|=0.1|N|$, $\sigma_p^2=10^{-3}$, $\beta_l=1$, $\beta_m = \beta_s=0.25$, and $|\Delta|$ was varied from 0 to $0.2|N|$. Also, $n=0.13m$ and $\sigma_w^2=10^{-5}$.
From the figure, we can observe the following. (1) Reg-mod-BPDN has much smaller bounds than those of mod-BPDN, BPDN and LS-CS. The differences between reg-mod-BPDN and mod-BPDN bounds is minor when $|\Delta|$ is small but increases as $|\Delta|$ increases. (2) The conclusions from the reconstruction error comparisons are similar to those seen from the bound comparisons, indicating that the bound can serve as a useful proxy to decide which algorithm to use when (notice bound computation is much faster than computing the reconstruction error). (3) Also, reg-mod-BPDN and mod-BPDN bounds are quite tight as compared to the LS-CS bound. BPDN bound and error are both $100\%$. 100\% error is seen because the reconstruction is the all zeros' vector.
In Fig. \ref{bound4comparefig}(c), we did a similar set of experiments for the case where $A$ corresponds to a simulated MRI experiment, i.e. $A=F_{u}\cdot W^{-1}$ where $F_u$ contains randomly selected rows of the 2D-DFT matrix and $W$ is the inverse 2D-DWT matrix (for a two-level Daubechies-4 wavelet). We used $n=0.17m$ and $\sigma_w^2=10^{-3}$. All other parameters were the same as in Fig. \ref{bound4comparefig}(b). Our conclusions are also the same.
The complexity for Theorem 3 is polynomial in $|\Delta|$ where as that of the LS-CS bound \cite[Corollary 1]{LSCSbound} is exponential in $|\Delta|$. To also show comparison with the LS-CS bound,
we had to choose a small value of $m=256$ so that the maximum value of $|\Delta|= 0.2|N|=5$ was small enough. In terms of MATLAB time, computation of the Theorem 3 bound for reg-mod-BPDN took 0.2 seconds while computing the LS-CS bound took 1.2 seconds.
For all methods except LS-CS, we were able to do the same thing fairly quickly even for $m=4096$, or even larger. It took only $8$ seconds to compute the bound of Theorem 3 when $m=4096$, $n=0.13m$, $|N|=410=0.1m$ and $|\Delta|=|\Delta_e|=0.1|N|=41$. The LS-CS bound, whose complexity is $O(2^{41})$ was taking too long (it did not get done even in a few minutes and we stopped the computation after that). The same was also true for the Theorem 2 bound which also has exponential complexity
{{
\section{Introduction}
The goal of this work is to solve the sparse recovery problem \cite{BPDN,justrelax,candescs,Donoho,rice}. We try to reconstruct an $m$-length sparse vector, $x$, with support, $N$, from an $n< m$ length noisy measurement vector, $y$, satisfying
\begin{equation}
y \triangleq Ax+w
\label{obsmod}
\end{equation}
when the following two things are available: (i) partial, and partly erroneous, knowledge of the signal's support, denoted by $T$; and (ii) an erroneous estimate of the signal values on $T$, denoted by $(\hat{\mu})_T$. In (\ref{obsmod}), $w$ is the measurement noise and $A$ is the measurement matrix.
For simplicity, in this work, we just refer to {\em $x$ as the signal} and to {\em $A$ as the measurement matrix}.
However, in general, $x$ is the sparsity basis vector (which is either the signal itself or some linear transform of the signal) and $A = H \Phi$ where $H$ is the measurement matrix
and $\Phi$ is the sparsity basis matrix. If $\Phi$ is the identity matrix, then $x$ is the signal itself
The true support of the signal, $N$, can be rewritten as
\begin{equation}
N = T \cup \Delta \setminus \Delta_e
\end{equation}
where
\begin{equation}
\Delta \triangleq N \setminus T \ \text{and} \ \Delta_e \triangleq T \setminus N
\label{defdelta}
\end{equation}
are the errors in the support estimate, $T^c$ is the complement set of $T$ and $ \setminus $ is the set difference notation ($N \setminus T \triangleq N \cap T^c $).
\newcommand{\muhat}{\hat\mu}
The signal estimate is assumed to be zero along $T^c$, i.e.
\begin{equation}
\hat{\mu} = \left[
\begin{array}{c}
(\hat{\mu})_T \\
\mathbf{0}_{T^c} \\
\end{array}
\right]
\label{sigestim}
\end{equation}
and the signal itself can be rewritten as
\begin{eqnarray}\label{signalesterror}
(x)_{N \cup T} &=& (\hat\mu)_{N \cup T} + e \\
(x)_{N^c} &=& 0
\end{eqnarray}
where $e$ denotes the error in the prior signal estimate. It is assumed that the error energy, $\|e\|_2^2$, is small compared to the signal energy, $\|x\|_2^2$.
In practical applications, $T$ and $\muhat$ may be available from prior knowledge. Alternatively, in applications requiring recursive reconstruction of (approximately) sparse signal or image sequences, with slow time-varying sparsity patterns and slow changing signal values, one can use the support estimate and the signal value estimate from the previous time instant as the ``prior knowledge". A key domain where this problem occurs is in fast (recursive) dynamic MRI reconstruction from highly undersampled measurements. In MRI, we typically assume that the images are wavelet sparse.
We show slow support and signal value change for two medical image sequences in Fig. \ref{slowchange}. From the figure, we can see the maximum support changes for both sequences are less than 2\% of the support size and almost all signal values' changes are less than $0.16\%$ of the signal energy. Slow signal value change also implies that a signal value is small before it gets removed from the support
This work has the following contributions.
\begin{enumerate}
\item We introduce regularized modified-BPDN (reg-mod-BPDN) and obtain a computable bound on its reconstruction error using the approach motivated by \cite{justrelax}. Reg-mod-BPDN solves
\begin{equation}\label{regmodBPDN0}
\min_b \ \ \gamma\|b_{T^c}\|_1 +\frac{1}{2}\|y-Ab\|_2^2 + \frac{1}{2}\lambda\|b_T-\hat{\mu}_T\|_2^2
\end{equation}
i.e. it tries to find the signal that is sparsest outside the set $T$, while being ``close enough" to $\hat{\mu}_T$ on $T$, and while satisfying the data constraint. Reg-mod-BPDN uses the fact that $T$ is a good estimate of the true support, $N$, and that $\muhat_T$ is a good estimate of $x_T$. In particular, for $i \in \Delta_e$, this implies that $|\muhat_i|$ is close to zero (since $x_i = 0$ for $i \in \Delta_e$).
\item Our second key contribution is to show how to use the reconstruction error bound result to obtain another computable bound that holds without any sufficient conditions and is tighter. This allows easy bound comparisons of the various approaches. A similar result for mod-BPDN and BPDN follows as a direct corollary.
\item Extensive reconstruction error comparisons with these and many other existing approaches are also shown
\end{enumerate}
\begin{figure*}
\centerline{
\subfigure[]{
\begin{tabular}{cc}
\epsfig{file = originallarynxseqremovewhitespace.eps,height=1.7cm}
\epsfig{file = noiselessFTsparsifyheart16obs_original_example.eps, height=1.7cm} \\%\epsfig{file = U:/KFCS/arxiv_lscs/vowelorg.eps,height=2cm} \vspace{-0.2in} \\
{\small (i) a larynx (vocal tract) image sequence} & {\small (ii) cardiac image sequence}
\end{tabular}
}
}
\centerline{
\subfigure[]{
\begin{tabular}{ccc}
\epsfig{file = slowchangedelta99.eps,height=3.5cm, width = 6.2cm} &
\epsfig{file = slowchangedeltae99.eps,height=3.5cm, width = 6.2cm} &
\epsfig{file = signalvaluechangecardiaclarynx99newylabel.eps,height=3.5cm, width = 6.7cm}\\
{\small (i) support additions, $\frac{|N_t \setminus N_{t-1}|}{|N_t|}$} & {\small (ii) support removals, $\frac{|N_{t-1} \setminus N_{t}|}{|N_t|}$} & {\small (iii) signal value change, $\frac{\|(x_t-x_{t-1})_{N_t\cup N_{t-1}}\|_2}{\|(x_t)_{N_t}\|_2}$}
\end{tabular}
}
}
\vspace{-0.1in}
\caption{{\small{In (a), we show two medical image sequences (a cardiac and a larynx sequence).
In (b), $x_t$ is the two-level Daubechies-4 2D discrete wavelet transform (DWT) of the cardiac or the larynx image at time $t$ and the set $N_t$ is its 99\% energy support (the smallest set containing 99\% of the vector's energy). Its size, $|N_t|$ varied between 4121-4183 ($\approx 0.07m$) for larynx and between 1108-1127 ($\approx 0.06m$) for cardiac. {\em Notice that all support changes are less than 2\% of the support size and almost all signal values changes are less than 4\% of $\|(x_t)_{N_t}\|_2$.}}}}
\label{slowchange}
\end{figure*}
\subsection{Notations and Problem Definition}
For any set $T$ and vector $b$, $b_T$ denotes a sub-vector
containing the elements of $b$ with indices in $T$.
$\|b\|_k$ refers to the $\ell_k$ norm of the vector $b$. Also, $\|b\|_0$ counts the number of nonzero elements of $b$.
The notation $T^c$ denotes the set complement of $T$, i.e., $ T^c = \{i \in [1,...,m], i \notin T \}$. $\emptyset$ is the empty set.
We use $'$ for transpose. For the matrix $A$, $A_T$ denotes the sub-matrix containing the columns of $A$ with indices in $T$. The matrix norm $\|A\|_{p}$, is defined as $\|A\|_{p} \triangleq \max_{x\neq 0}\frac{\|Ax\|_p}{\|x\|_p}$. $I_T$ is an identity matrix on the set of rows and columns indexed by elements in $T$. $\mathbf{0}_{T,S}$ is a zero matrix on the set of rows and columns indexed by elements in $T$ and $S$ respectively
The notation $\nabla L(b)$ denotes the gradient of function $L(b)$ with respect to $b$.
When we say {\em b is supported on $T \cup S$} we mean that the support of $b$ (set of indices where $b$ is nonzero) is a subset of $T \cup S$.
Our goal is to reconstruct a sparse vector, $x$, with support, $N$, from the noisy measurement vector, $y$ satisfying (\ref{obsmod}). We assume partial knowledge of the support, denoted by $T$, and of the signal estimate on $T$, denoted by $(\hat{\mu})_T$. The support estimate may contain errors -- misses, $\Delta$, and extras, $\Delta_e$, defined in (\ref{defdelta}). The signal estimate, $\muhat$, is assumed to be zero along $T^c$, i.e it satisfies (\ref{sigestim}) and the signal, $x$, satisfies (\ref{signalesterror})
\subsection{Related Work}
The sparse reconstruction problem, without using any support or signal value knowledge, has been studied for a long
time \cite{BPDN,justrelax,candescs,Donoho,rice}. It tries to find the sparsest signal among all signals that satisfy
the data constraint, i.e. it solves $\min_{b} \|b\|_0 \ s.t. \ y = A\beta$. This brute-force search has exponential complexity. One class of practical approaches to solve this involves replacing $\|b\|_0$ by $\|b\|_1$ which is the closest norm to $\ell_0$ that makes the problem convex (basis pursuit) \cite{BPDN}.
For noisy measurements, the data constraint becomes an inequality constraint. However, this assumes that the noise is bounded and the noise bound is available. In practical applications where this may not be available, one can use the Lagrangian version which solve
\begin{equation}\label{BPDN}
\min_b \ \gamma\|b\|_1 + \frac{1}{2}\|y-Ab\|_2^2
\end{equation}
This is called {\em basis pursuit denoising (BPDN) \cite{BPDN}}. Since this solves an unconstrained optimization problem, it is also faster (needed for large problems). The error bound of BPDN was obtained in \cite{justrelax}. Error bounds for the constrained version of BPDN were obtained in \cite{candes_rip, candesnoise}.
The problem of sparse reconstruction with partial support knowledge was introduced in our work \cite{modcsisit,modcsjournal};
and also in parallel in Khajehnejad et al \cite{weightedl1} and in vonBorries et al \cite{camsap07}.
In \cite{modcsisit,modcsjournal}, we proposed an approach called {\em modified-CS} which tries to find the signal that is
sparsest outside the set $T$ and satisfies the data constraint. We obtained exact reconstruction conditions for
it by using the restricted isometry approach \cite{decodinglp}. When measurements are noisy, for the same reasons as above, one can use the Lagrangian version:
\begin{equation}\label{mb}
\min_b \ \ \gamma\|b_{T^c}\|_1 + \frac{1}{2}\|y-Ab\|_2^2
\end{equation}
We call this {\em modified-BPDN (mod-BPDN)}. Its error was bounded in the conference version of this work \cite{modBPDN}, while the error of its constrained version was bounded in Jacques \cite{iBPDN}.
In \cite{weightedl1}, Khajehnejad et al assumed a probabilistic support prior and proposed a {weighted $\ell_1$} solution. They also obtained exact reconstruction thresholds for weighted $\ell_1$ by using the overall approach of Donoho \cite{Dohohowl1}. In Fig. \ref{reconfig}, we show comparisons with the noisy Lagrangian version of {\em weighted $\ell_1$} which solves
\begin{equation}
\min_b \ \ \gamma\|b_{T^c}\|_1+\gamma'\|b_T\|_1 + \frac{1}{2}\|y-Ab\|_2^2 \label{weightedl1}
\end{equation}
Our earlier work on Least Squares CS-residual (LS-CS) and Kalman Filtered CS-residual (KF-CS) \cite{LSCS,kfcsicip} can also be interpreted as a possible solution for the current problem, although it was proposed in the context of recursive reconstruction of sparse signal sequences.
Reg-mod-BPDN may also be interpreted as a Bayesian CS or a model-based CS approach. Recent work in this area includes \cite{modelCS, eldar_TIT_2009, bayesianCS, schniter,schniter_hmmtree, TOMP_do, TOMP_bniuk}.
\subsection{Some Related Approaches**}
Before going further, we discuss below {\em a few approaches that are related to, but different from reg-mod-BPDN, and we argue when and why these will be worse than reg-mod-BPDN.} This section may be skipped on a quick reading. We show comparisons with all these in Fig. \ref{reconfig}.
The first is what can be called {\em CS-residual or CS-diff} which computes
\begin{eqnarray}
&& \hat{x} = \hat{\mu} + \hat{b}, \ \text{where $\hat{b}$ solves} \nonumber \\
&& \min_{b} \ \ \gamma \|b\|_1 + \frac{1}{2}\|y-A\hat{\mu}-Ab\|_2^2
\label{CSresidual}
\end{eqnarray}
This has the following limitation. It does not use the fact that when $T$ is an accurate estimate of the true support, $(x)_{T^c}$ is much more sparse compared with the full $(x-\hat{\mu})$. The exception is if the signal value prior is so strong that $(x-\hat{\mu})$ is zero (or very small) on all or a part of $T$.
CS-residual is also related to LS-CS and KF-CS. LS-CS solves (\ref{CSresidual}) but with $\hat{\mu}_T$ being the LS estimate computed assuming that the signal is supported on $T$ and with $(\muhat)_{T^c}=0$. For a static problem, KF-CS can be interpreted as computing the regularized LS estimate on $T$ and using that as $\hat{\mu}_T$. LS-CS and KF-CS also have a limitation similar to CS-residual
Another seemingly related approach is what can be called {\em CS-mod-residual.} It computes
\begin{eqnarray}
&& \hat{x}_T = \hat{\mu}_T, \ \hat{x}_{T^c}=\hat{b}_c, \ \text{where $\hat{b}_c$ solves} \nonumber \\
&& \min_{b_c} \ \ \frac{1}{2}\|y-A_T\hat{\mu}_T-A_{T^c}b_c\|_2^2+\gamma \|b_c\|_1
\label{CSmodresidual}
\end{eqnarray}
where $b_c$ stands for $(b)_{T^c}$. This is solving a sparse recovery problem on $T^c$, i.e. it is implicitly assuming that $x_T$ is either equal to $\hat{\mu}_T$ or very close to it. Thus, this also works only when the signal value prior is very strong.
Both CS-residual and CS-mod-residual can be interpreted as extensions of BPDN, and \cite[Theorem 8]{justrelax} can be used to bound their error. In either case, the bound will contain terms proportional to $\|(x_T - \hat{\mu}_T)\|_2$ and as a result, it will be large when the prior is not good\footnote{In either case, one can assume that $(x-\hat\mu)$ is supported on $\Delta$ and the ``noise" is $w+A_T(x_T-\hat{\mu}_T)$. Thus, CS-residual error can be bounded by $C(A,\Delta)(\|w\|_2 +\|A_T(x_T-\hat{\mu}_T)\|_2)$ while CS-mod-residual error can be bounded by $\|x_T-\hat{\mu}_T\|_2 + C(A_{T^c},\Delta)(\|w\|_2 +\|A_T(x_T-\hat{\mu}_T)\|_2)$.}. This is also seen from our simulation experiments shown in Fig. \ref{reconfig} where we provide comparisons for the case of good signal value prior (0.1\% error in initial signal estimate) and bad signal value prior (10\% error in initial signal estimate). We vary support errors from 5\% to 20\% misses, while keeping the extras fixed at 10\%
Reg-mod-BPDN can also be confused with {\em modified-CS-residual} which computes\cite{modcsMRI}
\begin{eqnarray}
\hat{x} &=& \hat{\mu} + \hat{b}, \ \text{where $\hat{b}$ solves} \nonumber \\
&& \min_{b} \ \ \frac{1}{2}\|y-A\hat{\mu}-A b\|_2^2+\gamma \|b_{T^c}\|_1
\label{modCSresidual}
\end{eqnarray}
This is indeed related to reg-mod-BPDN and in fact this inspired it. We studied this empirically in \cite{modcsMRI}. However, one cannot get good error bounds for it in any easy fashion. Notice that the minimization is over the entire vector $b$, while the $\ell_1$ cost is only on $b_{T^c}$.
As pointed out by anonymous reviewer, one may consider solving the following variant of reg-mod-BPDN (we call this {\em reg-mod-BPDN-var}):
\begin{equation}\label{regmodBPDNvar}
\min_b \ \ \ \gamma\|b_{T^c}\|_1+\frac{1}{2}\|y-Ab\|_2^2+\frac{1}{2}\lambda\|b-\hat{\mu}\|_2^2
\end{equation}
Since $\muhat$ is supported on $T$, the regularization term can be rewritten as $\lambda\|b-\hat{\mu}\|_2^2 = \lambda\|b_T-\hat{\mu}_{T}\|_2^2+\lambda\|b_{T^c}\|_2^2$. Thus,
in addition to the $\ell_1$ norm cost on $b_{T^c}$ imposed by the first term, this last term is also imposing an $\ell_2$ norm cost on it. If $\lambda$ is large enough, the $\ell_2$ norm cost
will encourage the energy of the solution to be spread out on $T^c$, thus causing it to be less sparse. Since the true $x$ is very sparse on $T^c$ ($|\Delta|$ is small compared to the support
size also), we will end up with a larger recovery error\footnote{In the limit if $\sqrt{\lambda/2}$ is much larger than $\gamma$, we may get a completely non-sparse solution.}. We show an example
in Fig. \ref{regmodBPDNvarcomp}. Notice that the reg-mod-BPDN-var solution has many (about 7) more extras on $T^c$ than the reg-mod-BPDN solution and thus has a larger total error. Also, see Fig. \ref{reconfig}(a).
However, if we compare the two approaches for compressible signal sequences, e.g. the larynx sequence, it is difficult to say which will be better [see Fig. \ref{larynxcompareregmodBPDN}]
Finally, one may solve the following ({\em can call it reg-BPDN})
\begin{equation}\label{regmodBPDNvar_0}
\min_b \ \ \ \gamma\|b\|_1+\frac{1}{2}\|y-Ab\|_2^2+\frac{1}{2}\lambda\|b-\hat{\mu}\|_2^2
\end{equation}
This has two limitations. (1) Like CS-residual, this also does not use the fact that when $T$ is an accurate estimate of the true support, $(x)_{T^c}$ is much more sparse compared with the full $(x-\hat{\mu})$. (2) Its last term is the same as that of reg-mod-BPDN-var which also causes the same problem as above
\begin{figure*}[t!]
\center
\subfigure[Comparisons on $N$]{
\includegraphics [width=7.5cm,height=6cm]{regmodBPDNlambda0p002originalerror4andvarianterror12andCSreserror10partN.eps}
}
\subfigure[Comparisons on $N^c$]{
\includegraphics [width=7.5cm,height=6cm]{regmodBPDNlambda0p002originalerror4andvarianterror12andCSreserror10partNc.eps}
}
\vspace{-3mm}
\caption{\small{We plot the true 256-length signal, $x$, and its reconstructions using reg-mod-BPDN, reg-mod-BPDN-var and CS-residual.
In (a), the large leftmost elements are those on $N \cap T$ followed by the two smaller ones on $\Delta$. In (b),
we plot the reconstructions on $N^c$. The first two indices correspond to the set $\Delta_e$ and the rest are elements of $(N \cup T)^c$.
In (b), the original signal, $x$, is all zeros, where as the reconstructions have some error. We are showing one realization of the case of Fig. \ref{reconfig}(a).
}
} \label{regmodBPDNvarcomp}
\end{figure*}
\subsection{Paper Organization}
We introduce reg-mod-BPDN in Sec. II. We obtain computable bounds on its reconstruction error in Sec. III. The simultaneous comparison of upper bounds of multiple approaches becomes difficult because their results hold under different sufficient
conditions. In Sec. IV, we address this issue by showing how to obtain a tighter error bound that also holds without any
sufficient conditions and is still computable. In both sections, the bounds for mod-BPDN and BPDN follow as direct corollaries.
In Sec V, the above result is used for easy numerical comparisons between the upper bounds of various
approaches -- reg-mod-BPDN, mod-BPDN, BPDN and LS-CS and for numerically evaluating tightness of the bounds with both Gaussian measurements and partial Fourier measurements.
We also provide reconstruction error comparisons with CS-residual, LS-CS, KF-CS, CS-mod-residual, mod-CS-residual and reg-mod-BPDN-var, as well as with weighted $\ell_1$, mod-BPDN and BPDN for (a) static sparse recovery from random-Gaussian measurements; and for (b) recovering a larynx image sequence from simulated MRI measurements.
Conclusions are given in Sec. VI
\section{Regularized Modified-BPDN (Reg-mod-BPDN)}
Consider the sparse recovery problem when partial support knowledge is available. As explained earlier, one can solve mod-BPDN given in (\ref{mb}). When the support estimate is accurate, i.e. $|\Delta|$ and $|\Delta_e|$ are small, mod-BPDN provides accurate recovery with fewer measurements than what BPDN needs. However, it puts no cost on $b_T$ except the cost imposed by the data term. Thus, when very few measurements are available or when the noise is large, $b_T$ can become larger than required (in order
to reduce the data term). A similar, though lesser, bias will occur with weighted $\ell_1$ also when $\gamma' < \gamma$.
To address this, when reliable prior signal value knowledge is available, we can instead solv
\begin{equation}\label{regmodBPDN}
\min_b \ \ \ L(b)\triangleq \gamma\|b_{T^c}\|_1+\frac{1}{2}\|y-Ab\|_2^2+\frac{1}{2}\lambda\|b_T-\hat{\mu}_T\|_2^2
\end{equation}
which we call {\em reg-mod-BPDN}. Its solution, denoted by $\hat{x}$, serves as the reconstruction of the unknown signal, $x$. Notice that the first term helps to find the solution that is sparsest outside $T$, the second term imposes the data constraint while the third term imposes closeness to $\hat\mu$ along $T$.
Mod-BPDN is the special case of (\ref{regmodBPDN}) when $\lambda = 0$. BPDN is also a special case with $\lambda = 0$ and $T = \emptyset$ (so that $\Delta = N$).
\subsection{Limitations and Assumptions}
A key limitation of adding the regularizing term, $\lambda\|b_T-\hat{\mu}_T\|_2^2$ is as follows.
It encourages the solution to be close to $(\muhat)_{\Delta_e}$ which is not zero. As a result, $(\xhat)_{\Delta_e}$ will also not be zero (except if $\lambda$ is very small) even though $(x)_{\Delta_e}=0$. Thus, even in the noise-free case, reg-mod-BPDN will not achieve exact reconstruction. In both noise-free and noisy cases, if $(\muhat)_{\Delta_e}$ is large, $(\xhat)_{\Delta_e}$ being close to $(\muhat)_{\Delta_e}$ can result in large error. Thus, we need the assumption that $(\muhat)_{\Delta_e}$ is small
For the reason above, when we estimate the support of $\xhat$, we need to use a nonzero threshold, i.e. compute
\begin{eqnarray}
\hat{N} = \{i: |\xhat_i| > \rho \}
\end{eqnarray}
with a $\rho > 0$.
To get a small error reconstruction, reg-mod-BPDN requires the following:
\begin{enumerate}
\item $T$ is a good estimate of the true signal's support, $N$, i.e. $|\Delta|$ and $|\Delta_e|$ are small compared to $|N|$; and
\item $\muhat_T$ is a good estimate of $x_T$. For $i \in \Delta_e$, this implies that $|\muhat_i|$ is close to zero (since $x_i = 0$ for $i \in \Delta_e$).
\item
For accurate (exact) support estimation, we also need that most (all) nonzero elements of $x$ are larger than $\max_{i \in \Delta_e} |\muhat_i|$.
\end{enumerate}
The smallest nonzero elements of $x$ are usually on the set $\Delta$. In this case, the third assumption is equivalent to requiring that most (all) elements of $x_\Delta$ are larger than $\max_{i \in \Delta_e} |\muhat_i|$
\subsection{Dynamic Reg-Mod-BPDN for Recursive Recovery} \label{dynreg}
An important application of reg-mod-BPDN is for recursively reconstructing a time sequence of sparse signals from undersampled measurements, e.g. for dynamic MRI. To do this, at time $t$ we solve (\ref{regmodBPDN}) with $T=\Nhat_{t-1}$, where $\Nhat_{t-1}$ is the support estimate of the previous reconstruction, $\xhat_{t-1}$, and $(\muhat)_T = (\xhat_{t-1})_T$ and $(\muhat)_{T^c} = \mathbf{0}$. At the initial time, $t=0$, we can either initialize with BPDN, or with mod-BPDN using $T$ from prior knowledge, e.g. for wavelet sparse images, $T$ could be the set of indices of the approximation coefficients. We summarize the stepwise dynamic reg-mod-BPDN approach in Algorithm \ref{recursivealgo}. Notice that at $t=0$, one may need more measurements since the prior knowledge of $T$ may not be very accurate. Hence, we use $y_0 = A_0 x_0+w_0$ where $A_0$ is an $n_0 \times m$ measurement matrix with $n_0 > n$.
In Algorithm \ref{recursivealgo}, we should reiterate that for support estimation, we need to use a threshold $\rho > 0$. The threshold should be large enough so that most elements of $\Delta_{e,t}:=T \setminus N_{t} = \Nhat_{t-1} \setminus N_t$ do not get detected into the support.
To address an important concern of an anonymous reviewer, we briefly explain here why reg-mod-BPDN will be stable by describing our result from \cite{regmodBPDNstability}. By ``stable", we mean that the reconstruction error and the support estimation errors remain bounded by a time-invariant and small value at all times.
In \cite{regmodBPDNstability}, we showed the following for the constrained version of reg-mod-BPDN (but the same ideas apply to reg-mod-BPDN also).
If (i) $\rho$ is large enough (so that $\Nhat_t$ does not falsely detect any element that got removed from $N_t$);
(ii) the newly added elements to the current support, $N_t$, either get added at a large enough value to get detected immediately,
or within a finite delay their magnitude becomes large enough to get detected; and (iii) the matrix $A$ satisfies certain restricted
isometry property (RIP) \cite{decodinglp} conditions (for a given support size and support change size); then reg-mod-BPDN is stable.
\begin{algorithm}[h!]
\caption{{\bf \small Dynamic Reg-mod-BPDN}
At $t=0$, compute $\hat{x}_0$ as the solution of $\min_{b} \ \ \ \gamma\|(b)_{T^c}\|_1+\frac{1}{2}\|y_0-Ab\|_2^2$,
where $T$ is either empty or is available from prior knowledge.
Compute $\hat{N}_0 = \{i \in [1,...,m] : |(\hat{x}_{0})_i| > \alpha \}$. Set $T \leftarrow \hat{N}_0$ and
$(\muhat)_T \leftarrow (\hat{x}_{0})_T$
For $t>0$, do
\begin{enumerate}
\item {\em Reg-Mod-BPDN. } Let $T = \hat{N}_{t-1}$ and let $\hat\mu_T = (\hat{x}_{t-1})_T$.
Compute $\hat{x}_{t}$ as the solution of (\ref{regmodBPDN}).
\label{step1recursive}
\item {\em Estimate Support. } $\hat{N}_t=\{i \in [1,...,m] : |(\hat{x}_{t})|_i > \rho \}$.
\item Output the reconstruction $\hat{x}_{t}$.
\end{enumerate}
Feedback $\Nhat_t$ and $\hat{x}_t$; increment $t$, and go to step \ref{step1recursive}.
\label{recursivealgo}
\end{algorithm}
\section{Bounding the Reconstruction Error}
In this section, we bound the reconstruction error of reg-mod-BPDN. Since mod-BPDN and BPDN are special cases, their results follow as direct corollaries. The result for BPDN is the same as \cite[Theorem 8]{justrelax}. In Sec. III-A, we define the terms needed to state our result. In III-B we state our result and discuss its implications. In III-C, we give the proof outline.
\subsection{Definitions}
We begin by defining the function that we want to minimize as
\begin{equation}
L(b) \triangleq L_1(b)+\gamma\|b_{T^c}\|_1 \label{regmodBPDNfunction}
\end{equation}
where
\begin{equation}
L_1(b) \triangleq \frac{1}{2}\|y-Ab\|_2^2+\frac{1}{2}\lambda\|b_T-\hat{\mu}_T\|_2^2
\end{equation}
contains the two $\ell_2$ norm terms (data fidelity term and the regularization term). If we constrain $b$ to be supported on $T \cup S$ for some $S \subset T^c$, then the minimizer of $L_1(b)$ will be the regularized least squares (LS) estimator obtained when we put a weight $\lambda$ on $\|b_T-\hat{\mu}_T\|_2^2$ and a weight zero on $\|b_S-\hat{\mu}_S\|_2^2$.
Let $S$ be a given subset of $\Delta$. Next, we define three matrices which will be frequently used in our results. Let
\begin{eqnarray}
Q_{T,\lambda}(S) & \triangleq & {A_{T\cup S}}'A_{T\cup S}+\lambda \left[
\begin{array}{cc}
I_{T}\ & \mathbf{0}_{T,S} \\
\mathbf{0}_{S,T}\ & \mathbf{0}_{S,S} \\
\end{array}
\right] \label{def_Q} \\
M_{T,\lambda} & \triangleq & I-A_T({A_T}'A_T+\lambda I_{T})^{-1}{A_T}' \\
P_{T,\lambda}(S)& \triangleq & ({A_{S}}'M_{T,\lambda}A_{S})^{-1} \label{def_P}
\end{eqnarray}
where $I_T$ is a $|T|\times |T|$ identity matrix and $\mathbf{0}_{T,S}$, $\mathbf{0}_{S,T}$, $\mathbf{0}_{S,S}$ are
all zeros matrices with sizes $|T|\times |S|$, $|S|\times |T|$ and $|S|\times |S|$.
\begin{assum}
We assume that $Q_{T,\lambda}(\Delta)$ is invertible. This implies that, for any $S \subseteq \Delta$, the functions $L(b)$ and $L_1(b)$ are strictly convex over the set of all vectors supported on $T \cup S$
\label{assum1}
\end{assum}
\begin{proposition}
When $\lambda > 0$, $Q_{T,\lambda}(S)$ is invertible if $A_S$ has full rank. When $\lambda =0$ (mod-BPDN), this will hold if $A_{T \cup S}$ has full rank.
\end{proposition}
The proof is easy and is given in Appendix \ref{Appprop1}.
Let $S \subseteq \Delta$. Consider minimizing $L(b)$ over $b$ supported on $T \cup S$. When $b_{(T \cup S)^c} = 0$ and Assumption \ref{assum1} holds, $L(b_{T \cup S})$ is strictly convex and thus has a unique minimizer. The same holds for $L_1(b_{T \cup S})$. Define their respective unique minimizers a
\begin{eqnarray}
d_{T,\lambda}(S) &\triangleq& \arg \min_b \ L(b) \ \ \text{ subject to } \ \ b_{(T\cup S)^c}=\mathbf{0} \ \ \label{regminimizer} \\
c_{T,\lambda}(S) &\triangleq& \arg \min_b \ L_1(b) \ \ \text{ subject to } \ \ b_{(T\cup S)^c}=\mathbf{0} \ \
\end{eqnarray}
As explained earlier, $c_{T,\lambda}(S)$ is the regularized LS estimate of $x$ when assuming that $x$ is supported on $T \cup S$ and with the weights mentioned earlier. It is easy to see that
\begin{eqnarray}\label{reglssolution}
\ [c_{T,\lambda}(S)]_{T\cup S} &=&Q_{T,\lambda}(S)^{-1} \left({A_{T\cup S}}'y + \left[
\begin{array}{c}
\lambda \hat{\mu}_{T} \\
\mathbf{0}_{S} \\
\end{array}
\right] \right) \nonumber \\
\ [c_{T,\lambda}(S)]_{(T\cup S)^c} &=& \mathbf{0} \label{def_c}
\end{eqnarray}
In a fashion similar to \cite{justrelax}, defin
\begin{eqnarray}
ERC_{T,\lambda}(S)& \triangleq & 1-\max_{\omega \notin
T\cup S}\|P_{T,\lambda}(S){A_{S}}'M_{T,\lambda}A_{\omega}\|_1
\end{eqnarray}
This is different from the ERC of \cite{justrelax} but simplifies to it when $T=\emptyset$, $S=N$ and $\lambda=0$. In \cite{justrelax}, the ERC,
which in our notation is $ERC_{\emptyset,0}(N)$, being strictly positive, along with $\gamma$ approaching zero, ensured exact recovery of BPDN
in the noise-free case. Hence, in \cite{justrelax}, ERC was an acronym for {\em Exact Recovery Coefficient}. In this work, the same holds for
mod-BPDN. If $ERC_{T,0}(\Delta)>0$, the solution of mod-BPDN approaches the true $x$ as $\gamma$ approaches zero.
We explain this further in Remark \ref{erc_modcs} below. However, no similar claim can be made for reg-mod-BPDN. On the other hand,
for the reconstruction error bounds, ERC serves the exact same purpose for reg-mod-BPDN as it does for BPDN in \cite{justrelax}:
$ERC_{T,\lambda}(\Delta) > 0$ and $\gamma$ greater than a certain lower bound ensures that the reg-mod-BPDN (or mod-BPDN) error can be bounded by modifying the approach of \cite{justrelax}.
\subsection{Reconstruction error bound}
The reconstruction error can be bounded as follows.
\begin{theorem}
If $Q_{T,\lambda}(\Delta)$ is invertible, $ERC_{T,\lambda}(\Delta) > 0$ and
\begin{equation}
\gamma \ge \gamma^*_{T,\lambda}(\Delta) \triangleq \frac{\|{A_{(T\cup \Delta)^c}}'(y-Ac_{T,\lambda}(\Delta))\|_{\infty}}{ERC_{T,\lambda}(\Delta)} \label{regmodBPDNcond}
\end{equation}
then,
\begin{enumerate}
\item $L(b)$ has a unique minimizer, $\hat{x}$.
\item The minimizer, $\hat{x}$, is equal to $d_{T,\lambda}(\Delta)$, and thus is supported on $T\cup \Delta$.
\item Its error can be bounded as
\begin{eqnarray}
& & \|x-\hat{x}\|_2 \le \gamma \sqrt{|\Delta|} f_1(\Delta)+\lambda f_2(\Delta)\|x_T-\hat\mu_T\|_2\nonumber \\
& & \hspace{18mm} +f_3(\Delta)\|w\|_2 \nonumber\\
& &\text{where } \nonumber \\
& & f_1(\Delta) \triangleq \nonumber \\
& & \sqrt{\|({A_T}'A_T + \lambda I_{T})^{-1} {A_T}'A_{\Delta}P_{T,\lambda}(\Delta)\|_{2}^2 + \|P_{T,\lambda}(\Delta)\|_{2}^2}, \nonumber\\
& & f_2(\Delta) \triangleq \|Q_{T,\lambda}(\Delta)^{-1}\|_2, \nonumber\\
& & f_3(\Delta) \triangleq \|Q_{T,\lambda}(\Delta)^{-1}{A_{T\cup \Delta}}'\|_2, \ \hspace{-8mm} \label{loosebound}
\end{eqnarray}
$P_{T,\lambda}(\Delta)$ is defined in (\ref{def_P}) and $Q_{T,\lambda}(\Delta)$ in (\ref{def_Q}).
\end{enumerate}
\label{thm1}
\end{theorem}
\begin{corollary}[corollaries for mod-BPDN and BPDN]
The result for mod-BPDN follows by setting $\lambda=0$ in Theorem \ref{thm1}. The result for BPDN follows by setting $\lambda=0$, $T=\emptyset$ (and so $\Delta=N$). This result is the same as \cite[Theorem 8]{justrelax}.
\end{corollary}
\begin{remark}[smallest $\gamma$]
Notice that the error bound above is an increasing function of $\gamma$. Thus $\gamma=\gamma^*_{T,\lambda}(\Delta)$ gives the smallest bound
\end{remark}
In words, Theorem 1 says that, if $Q_{T,\lambda}(\Delta)$ is invertible, $ERC_{T,\lambda}(\Delta)$ is positive, and $\gamma$ is large
enough (larger than $\gamma^*$), then $L(b)$ has a unique minimizer, $\xhat$, and $\xhat$ is supported on $T \cup \Delta = N \cup \Delta_e$.
This means that the only wrong elements that can possibly be part of the support of $\xhat$ are elements of $\Delta_e$.
Moreover, the error between $\xhat$ and the true $x$ is bounded by a value that is small as long as the noise, $\|w\|_2$, is small,
the prior term, $\|x_T- \muhat_T\|_2$, is small and $\gamma^*_{T,\lambda}(\Delta)$ is small. By rewriting $y - Ac_{T,\lambda}(\Delta) = A(x - c_{T,\lambda}(\Delta)) + w$ and using Lemma 2 (given in the Appendix)
one can upper bound $\gamma^*$ by terms that depend on $\|w\|_2$ and $\|x_T- \muhat_T\|_2$. Thus, as long as these two are small,
the bound is small
As shown in Proposition 1, $Q_{T,\lambda}(\Delta)$ is invertible if $\lambda>0$ and $A_\Delta$ is full rank or if $A_{T \cup \Delta}$ is full rank.
Next, we use the idea of \cite[Corollary 10]{justrelax} to show that $ERC_{T,0}(\Delta)$ is an {\em Exact Recovery Coefficient} for mod-BPDN
\begin{remark}[ERC and exact recovery of mod-BPDN]
For mod-BPDN, $c_{T,0}(\Delta)$ is the LS estimate when $x$ is supported on $T \cup \Delta$. Using (\ref{def_c}), (\ref{obsmod}), and the fact that $x$ is supported on $N \subseteq T \cup \Delta$, it is easy to see that in the noise-free ($w=0$) case, $c_{T,0}(\Delta) = x_{T \cup \Delta}$. Hence the numerator of $\gamma^*_{T,0}(\Delta)$ will be zero. Thus, using Theorem 1, if $ERC_{T,0}(\Delta)>0$, the mod-BPDN error satisfies $\|x- \xhat\|_2 \le \gamma \sqrt{|\Delta|} f_1(\Delta)$. Thus the mod-BPDN solution, $\xhat$, will approach the true $x$ as $\gamma$ approaches zero.
Moreover, as long as $\gamma < \frac{\min_{i \in N}|x_i|}{\sqrt{|\Delta|} f_1(\Delta)}$, at least the support of $\xhat$ will equal the true support, $N$ \footnote{If we bounded the $\ell_\infty$ norm of the error as done in \cite{justrelax} we would get a looser upper bound on the allowed $\gamma$'s for this.}.
\label{erc_modcs}
\end{remark}
We show a numerical comparison of the results of reg-mod-BPDN, mod-BPDN and BPDN in Table I (simulation details given in Sec. V).
Notice that BPDN needs $90\%$ measurements for its sufficient conditions to start holding (ERC to become positive) where as mod-BPDN only needs $19\%$.
Moreover, even with $90\%$ measurements, the ERC of BPDN is just positive and very small. As a result, its error bound is
large ($27\%$ normalized mean squared error (NMSE)). Similarly, notice that mod-BPDN needs $n \ge 19\%$ for its sufficient conditions to start holding ($A_{T \cup \Delta}$ to become full rank which is needed for $Q_{T,0}(\Delta)$ to be invertible). For reg-mod-BPDN which only needs $A_\Delta$ to be full rank, $n=13\%$ suffices
\begin{remark}
A sufficient conditions' comparison only provides a comparison of when a given result can be applied to provide a bound on the reconstruction error. In other words, it tells us under what conditions we can guarantee that the reconstruction error of a given approach will be small (below a bound). Of course this does not mean that we cannot get small error even when the sufficient condition does not hold, e.g., in simulations, BPDN provides a good reconstruction using much lesser than 90\% measurements. However, when $n<90\%$ we cannot bound its reconstruction error using Theorem 1 above.
\end{remark}
Due to lack of space, we have removed the discussion relating the ERC to the RIP constants and then using the resulting lower bound for a sufficient conditions' comparison. We have moved it to the Supplementary Material.
\subsection{Proof Outline}
To prove Theorem \ref{thm1}, we use the following approach motivated by that of \cite{justrelax}.
\begin{enumerate}
\item We first bound $\|d_{T,\lambda}(\Delta)-c_{T,\lambda}(\Delta)\|_2$ by simplifying the necessary and sufficient condition for it to be the minimizer of $L(b)$ when $b$ is supported on $T \cup \Delta$.
This is done in Lemma 1 in Appendix \ref{App3lemmas}.
\item We bound $\|c_{T,\lambda}(\Delta)-x\|_2$ using the expression for $c_{T,\lambda}(\Delta)$ in (\ref{reglssolution}) and substituting $y = A_{T \cup \Delta} x_{T \cup \Delta} + w$ in it (recall that $x$ is zero outside $T \cup \Delta$). This is done in Lemma 2 in Appendix \ref{App3lemmas}.
\item We can bound $\|d_{T,\lambda}(\Delta)-x\|_2$ using the above two bounds and the triangle inequality.
\item We use an approach similar to \cite[Lemma 6]{justrelax} to find the sufficient conditions under which $d_{T,\lambda}(\Delta)$ is also the unconstrained unique minimizer of $L(b)$, i.e. $\hat{x} = d_{T,\lambda}(\Delta)$. This is done in Lemma 3 in Appendix \ref{App3lemmas}
\end{enumerate}
The last step (Lemma 3) helps prove the first two parts of Theorem 1. Combining the above four steps, we get the third part (error bound).
We give the lemmas in Appendix \ref{App3lemmas}. They are proved in Appendix \ref{Applemma1proof}, \ref{Applemma2proof} and \ref{Applemma3proof}
Two key differences in the above approach with respect to the result of \cite{justrelax} are
\begin{itemize}
\item $c_{T,\lambda}(\Delta)$ is the regularized LS estimate instead of the LS estimate in \cite{justrelax}.
This helps obtain a better and simpler error bound of reg-mod-BPDN than when using the LS estimate. Of course, when $\lambda=0$ (mod-BPDN or BPDN),
$c_{T,0}(\Delta)$ is just the LS estimate again.
\item For reg-mod-BPDN (and also for mod-BPDN), the subgradient set of the $\ell_1$ term
is $\partial \|b_{T^c}\|_1|_{b=d_{T,\lambda}(\Delta)}$ and so any $\phi$ in this set is zero on $T$,
and only has $\|\phi_{\Delta}\|_{\infty}\le 1$. Since $|\Delta| \ll |N|$, this helps to get a tighter bound
on $\|c_{T,\lambda}(\Delta) - d_{T,\lambda}(\Delta)\|_2$ in step 1 above as compared to that for
BPDN \cite{justrelax} (see proof of Lemma 1 for details).
\end{itemize}
\section{Tighter Bounds without Sufficient Conditions}
The problem with the error bounds for reg-mod-BPDN, mod-BPDN, BPDN or LS-CS \cite{LSCSbound} is that they all hold under different sufficient conditions. This makes it difficult to compare them. Moreover, the bound is particularly loose when $n$ is such that the sufficient conditions just get satisfied. This is because the ERC is just positive but very small (resulting in a very large $\gamma^*$ and hence a very large bound).
To address this issue, in this section, we obtain a bound that holds without any sufficient conditions and that is also tighter, while still being computable.
The key idea that we use is as follows:
\begin{itemize}
\item we modify Theorem 1 to hold for ``sparse-compressible" signals \cite{LSCSbound}, i.e. for sparse signals, $x$, in which some nonzero coefficients out of the set $\Delta$ are small (``compressible") compared to the rest; and then
\item we minimize the resulting bound over all allowed split-ups of $x$ into non-compressible and compressible parts
\end{itemize}
Let $\tilde{\Delta} \subseteq \Delta$ be such that the conditions of Theorem 1 hold for it. Then the first step involves modifying Theorem 1 to bound the error for reconstructing $x$ when we treat $x_{\Delta \setminus \tilde{\Delta}}$ as the ``compressible" part. The main difference here is in bounding $\|c_{T,\lambda}(\tilde{\Delta})-x\|_2$ which now has a larger bound because of $x_{\Delta \setminus \tilde{\Delta}}$. We do this in Lemma \ref{lemma4} in the Appendix \ref{Applemma4}. Notice from the proofs of Lemma 1 and Lemma 3 in Appendix \ref{Applemma1proof} and \ref{Applemma3proof} that nothing in their result changes if we replace $\Delta$ by a $\tDelta \subseteq \Delta$. Combining Lemma 4 with Lemmas 1 and 3 applied for $\tDelta$ instead of $\Delta$ leads to the following corollary.
\begin{corollary}
Consider a $\tDelta \subseteq \Delta$. If $Q_{T,\lambda}(\tilde{\Delta})$ is invertible, $ERC_{T,\lambda}(\tilde{\Delta})>0$, and $\gamma=\gamma^*_{T,\lambda}(\tilde{\Delta})$, then
\begin{equation}
\|x-\hat{x}\|_2 \le f(T,\lambda,\Delta,\tilde{\Delta},\gamma^*_{T,\lambda}(\tilde{\Delta}))
\end{equation}
where
\begin{eqnarray}
f(T,\lambda,\Delta,\tilde{\Delta},\gamma) & \triangleq& \gamma \sqrt{|\tilde{\Delta}|} f_1(\tDelta)+ \lambda f_2(\tDelta) \|x_{T}-\hat{\mu}_{T}\|_2 \nonumber\\ & & + f_3(\tDelta)\|w\|_2 +
f_4(\tDelta) \|x_{\Delta \setminus \tilde{\Delta}}\|_2, \label{def_f} \\
f_4(\tDelta) & \triangleq & \sqrt{\|Q_{T,\lambda}(\tilde{\Delta})^{-1}{A_{T\cup \tilde{\Delta}}}'A_{\Delta \setminus \tilde{\Delta}}\|_2^2+1}, \ \ \ \ \
\label{def_f4}
\end{eqnarray}
$f_1(\cdot)$,$f_2(\cdot)$, $f_3(\cdot)$ are defined in (\ref{loosebound}) and $\gamma^*_{T,\lambda}(\tilde{\Delta})$ in (\ref{regmodBPDNcond}).
\label{cor2}
\end{corollary}
{\em Proof: } The proof is given in Appendix \ref{Appcor1proof}.
In order to get a bound that depends only on $\|x_T-\hat{\mu}_T\|_2$, $\|x_{\Delta \setminus \tDelta}\|_2$, the noise, $w$, and the sets $T,\Delta,\Delta_e$, we can further bound $\gamma^*_{T,\lambda}(\tilde{\Delta})$ by rewriting $y-Ac_{T,\lambda}(\tilde{\Delta}) = A(x-c_{T,\lambda}(\tilde{\Delta})) + w$ and then bounding $\|x-(c_{T,\lambda}(\tilde{\Delta}))\|_2$ using Lemma 4. Doing this gives the following corollary.
\begin{corollary}
If $Q_{T,\lambda}(\tilde{\Delta})$ is invertible, $ERC_{T,\lambda}(\tilde{\Delta})>0$, and $\gamma=\gamma^*_{T,\lambda}(\tilde{\Delta})$, then
\begin{equation}
\|x-\hat{x}\|_2 \le g(\tDelta) \label{gdefbound}
\end{equation}
where
\begin{eqnarray}
\hspace{-10mm} g(\tDelta) & \defn & g_1 \|x_T - \hat{\mu}_T\|_2 + g_2 \|w\|_2 + g_3 \|x_{\Delta \setminus \tDelta}\|_2+g_4 \hspace{7mm} \label{gbound} \\
g_1 & \defn & \lambda f_2(\tDelta) (\frac{\sqrt{|\tDelta|}f_1(\tDelta) \text{maxcor}(\tDelta)}{ERC_{T,\lambda}(\tDelta)}+1 ), \nonumber \\
g_2 & \defn & \frac{\sqrt{|\tDelta|}f_1(\tDelta)f_3(\tDelta) \text{maxcor}(\tDelta) }{ERC_{T,\lambda}(\tDelta)}+f_3(\tDelta), \nonumber \\
g_3 & \defn & \frac{\sqrt{|\tDelta|}f_1(\tDelta)f_4(\tDelta) \text{maxcor}(\tDelta) }{ERC_{T,\lambda}(\tDelta)}+f_4(\tDelta), \nonumber \\
g_4 & \defn & \frac{\sqrt{|\tDelta|}\|{A_{(T\cup \tDelta)^c}}'w\|_{\infty}f_1(\tDelta)}{ERC_{T,\lambda}(\tDelta)}, \quad \nonumber\\
\text{maxcor}(\tDelta)& \defn & \max_{i\notin (T\cup \tDelta)^c}\|{A_i}'A_{T\cup \Delta}\|_2, \nonumber
\end{eqnarray}
$f_1(\cdot)$,$f_2(\cdot)$, $f_3(\cdot)$ and $f_4(\cdot)$ are defined in (\ref{loosebound}) and (\ref{def_f4}), and $\gamma^*_{T,\lambda}(\tilde{\Delta})$ in (\ref{regmodBPDNcond}).
\label{corol2}
\end{corollary}
{\em Proof: } The proof is given in Appendix \ref{Appcor2proof}.
Using the above corollary and minimizing over all allowed $\tDelta$'s, we get the following result.
\begin{theorem}
Let
\begin{equation}
\tilde{\Delta}^* \triangleq \ \ \arg \min_{\hspace{-8mm}\tilde{\Delta}\in \mathcal{G}} g(\tDelta)
\end{equation}
where
\begin{equation}
\mathcal{G} \triangleq \{\tilde{\Delta}:\tilde{\Delta}\subseteq \Delta, ERC_{T,\lambda}(\tilde{\Delta})>0, Q_{T,\lambda}(\tilde{\Delta}) \text{ is invertible}\}
\end{equation}
If $\gamma=\gamma^*_{T,\lambda}(\tilde{\Delta}^*)$, then
\begin{enumerate}
\item $L(b)$ has a unique minimizer, $\hat{x}$, supported on $T\cup \tilde{\Delta}^*$.
\item The error bound is
\begin{equation}
\|x-\hat{x}\|_2\le g(\tDelta^*)
\label{regmodBPDNtightbound}
\end{equation}
\end{enumerate}
($\gamma^*_{T,\lambda}(\tilde{\Delta})$ is defined in (\ref{regmodBPDNcond})).
\label{thm2}
\end{theorem}
{\em Proof: } This result follows by minimizing over all allowed $\tDelta$'s from Corollary \ref{corol2}.
Compare Theorem 2 with Theorem 1. Theorem 1 holds only when the complete set $\Delta$ belongs to $\mathcal{G}$, where as Theorem 2 holds always (we only need to set $\gamma$ appropriately). Moreover, even when $\Delta$ does belong to $\mathcal{G}$, Theorem 1 gives the error bound by choosing $\tilde{\Delta}^* = \Delta$. However, the above result minimizes over all allowed ${\tDelta}$'s, thus giving a tighter bound, especially for the case when the sufficient conditions of Theorem 1 just get satisfied and $ERC_{T,\lambda}(\Delta)$ is positive but very small. A similar comparison also holds for the mod-BPDN and BPDN results.
The problem with Theorem 2 is that its bound is not computable (the computational cost is exponential in $|\Delta|$). Notice that $g(\tDelta^*):=\min_{\tilde{\Delta}\in \mathcal{G}} g(\tDelta)$ can be rewritten as
\begin{eqnarray}
& & g(\tDelta^*)\triangleq \min_{\tilde{\Delta} \in \mathcal{G}} g(\tDelta) = \min_{0 \le k \le |\Delta|} \min_{\mathcal{G}_k} g(\tDelta) \ \ \text{where} \nonumber \\
& & \mathcal{G}_k \triangleq \mathcal{G} \cap \{\tDelta \subseteq \Delta : |\tDelta| = k\}
\label{Gkdef}
\end{eqnarray}
Let $d:=|\Delta|$. The minimization over $\mathcal{G}_k$ is expensive since it requires searching over all ${d \choose k}$ size $k$ subsets of $\Delta$ to first find which ones belong to $\mathcal{G}_k$ and then find the minimum over all $\tDelta \subseteq \mathcal{G}_k$. The total computation cost to do the former for all sets $\mathcal{G}_0, \mathcal{G}_1, \dots \mathcal{G}_d$ is $O(\sum_{k=0}^d {d \choose k}) = O(2^d)$, i.e. it is {\em exponential in $d$}. This makes the bound computation intractable for large problems.
\subsection{Obtaining a Computable Bound}
In most cases of practical interest, the term that has the maximum variability over different sets in $\mathcal{G}_k$ is $\|x_{\Delta \setminus \tDelta}\|_2$. The multipliers $g_1$, $g_2$, $g_3$ and $g_4$ vary very slightly for different sets in a given $\mathcal{G}_k$. Using this fact, we can obtain the following upper bound on $\min_{\mathcal{G}_k} g(\tDelta)$ which is only slightly looser and also holds without sufficient conditions, but is computable in polynomial time.
Define $\tDelta^{**}(k)$ and $B_k$ as follows
\begin{eqnarray}
\tDelta^{**}(k) & \defn & \arg \min_{ \{\tDelta \subseteq \Delta, |\tDelta|=k \} } \|x_{\Delta \setminus \tDelta}\|_2 \nonumber \\
B_k & \defn & \left\{ \begin{array}{cc}
g(\tDelta^{**}(k)) & \ \text{if} \ \tDelta^{**}(k) \in \mathcal{G}_k \\
\infty & \ \text{otherwise} \
\end{array}
\right.
\label{defBk}
\end{eqnarray}
Then, clearly
\begin{eqnarray}
\min_{\mathcal{G}_k} g(\tDelta) \le B_k
\label{Bk_bnd}
\end{eqnarray}
since $\min_{\mathcal{G}_k} g(\tDelta) \le g(\tDelta)$ for any $\tDelta \in \mathcal{G}_k$ and it is also less than infinity.
For any $k$, the set $\tDelta^{**}(k)$ can be obtained by sorting the elements of $x_\Delta$ in decreasing order of magnitude and letting $\tDelta^{**}(k)$ contain the indices of the $k$ largest elements. Doing this takes $O(d \log d)$ time since sorting takes $O(d \log d)$ time. Computation of $B_k$ requires matrix multiplications and inversions which are $O(k^3)$. Thus, the total cost of doing this is at most $O(d^4)$ which is still polynomial in $d$.
Therefore, we get the following bound that is {\em computable in polynomial time and that still holds without
sufficient conditions and is much tighter than Theorem 1}.
\begin{theorem}
Let
\begin{eqnarray}
k_{\min} & \defn & \arg \min_{0 \le k \le |\Delta|} B_k \quad \text{ and }\label{kmindef} \nonumber \\
\tDelta^{**} & \defn & \tDelta^{**}(k_{\min})
\end{eqnarray}
where $B_k$ and $\tDelta^{**}(k)$ are defined in (\ref{defBk}).
If $\gamma=\gamma^*_{T,\lambda}(\tDelta^{**})$
\begin{enumerate}
\item $L(b)$ has a unique minimizer, $\hat{x}$, supported on $T\cup \tDelta^{**}$.
\item The error bound is
\begin{equation}
\|x-\hat{x}\|_2 \le g(\tDelta^{**})
\label{regmodBPDNmediumtightbound}
\end{equation}
\end{enumerate}
($\gamma^*_{T,\lambda}(\tilde{\Delta})$ is defined in (\ref{regmodBPDNcond})).
\label{thm3}
\end{theorem}
\begin{corollary}[corollaries for mod-BPDN and BPDN]
The result for mod-BPDN follows by setting $\lambda=0$ in Theorem \ref{thm3}. The result for BPDN follows by setting $\lambda=0$, $T=\emptyset$ (and so $\Delta=N$) in Theorem \ref{thm3}.
\end{corollary}
When $n$ and $s \triangleq |N|$ are large enough, the above bound is either only slightly larger, or often actually equal, to that of Theorem 2 (e.g. in Fig. \ref{bound4comparefig}(a), $m=256$, $n =0.13m=33$, $s=0.1m=26$). The reason for the equality is that the minimizing value of $k$ is the one that is small enough to ensure that $g_1, g_2, g_3,g_4$ are small. When $k$ is small, $g_1, g_2, g_3, g_4$, ${ERC}$ and $Q(\tDelta)$ have very similar values for all sets $\tDelta$ of the same size $k$. In (\ref{gbound}), the only term with significant variability for different sets $\tDelta$ of the same size $k$ is $\|x_{\Delta \setminus \tDelta}\|_2$. Thus,
(a) $\arg\min_{\mathcal{G}_k} g(\tDelta) = \arg\min_{\mathcal{G}_k} \|x_{\Delta \setminus \tDelta}\|_2$ and
(b) $\mathcal{G}_k$ is equal to $\{\tDelta \subseteq \Delta, |\tDelta|=k\}$.
Thus, (\ref{Bk_bnd}) holds with equality and so the bounds from Theorems 3 and 2 are equal.
As $n$ and $s\triangleq |N|$ approach infinity, {\em one can, in fact, use a law of large numbers (LLN) argument to prove that both bounds will be equal with high probability (w.h.p.)}. The key idea here is the same as above: we show that as $n,s$ go to infinity, w.h.p., $g_1,g_2,g_3,g_4$, $Q$ and $ERC$ are equal for all sets $\tDelta$ of any given size $k$. The main idea of this argument is given in the Supplementary Material. It will be developed in future work
\input{expts_3}
\section{Conclusions and Future Work}
In this work we studied the problem of sparse reconstruction from noisy undersampled measurements when partial and partly erroneous, knowledge of the signal's support and an erroneous estimate of the signal values on the ``partly known support" is also available. Denote the support knowledge by $T$ and the signal value estimate on $T$ by $\hat{\mu}_T$. We proposed and studied a solution called regularized modified-BPDN which tries to find the signal that is sparsest outside $T$, while being ``close enough" to $\hat{\mu}_T$ on $T$, and while satisfying the data constraint. We showed how to obtain computable error bounds that hold without any sufficient conditions. This made it easy to compare bounds for the various approaches (corresponding results for modified-BPDN and BPDN follow as direct corollaries). Exhaustive empirical error comparisons with these and many other existing approaches are also provided.
In ongoing work, we are evaluating the utility of reg-mod-BPDN for recursive functional MR imaging to detect brain activation patterns in response to stimuli. On the other end, we are also working on obtaining conditions under which it will remain ``stable" (its error will be bounded by a time-invariant and small value) for a recursive recovery problem \cite{regmodBPDNstability}.
|
\section{Introduction}\label{Sec1}
Recently, Verlinde \cite{Verlinde} presented a remarkable new idea that gravity
can be explained as an entropic force caused by the information changes when a
material body moves away from the holographic screen. This idea implies that
gravity is not fundamental. With the holographic principle and the
equipartition law of energy, Verlinde showed that Newton's law of gravitation
can arise naturally and unavoidably in a theory in which space is emergent
through a holographic scenario, and a relativistic generalization leads to the
Einstein equations. In fact, the similar idea can be traced back to Sakharov's
work \cite{Sakharov}. On the other side, using the equipartition law of energy
for the horizon degrees of freedom together with the thermodynamic relation
$S=E/(2T)$, Padmanabhan also obtained the Newton's law of gravity
\cite{Padmanabhan1,Padmanabhan2}.
Subsequently, with the idea of entropic force, some applications have
been carried out. The Friemdann equations and the modified Friedmann
equations for Friedmann-Robertson-Walker universe in Einstein gravity
\cite{Y.G.Gong1,R.G.Cai}, $f(R)$ gravity \cite{Y.G.Gong2}, deformed
Horava-Lifshitz gravity \cite{WeiLiu1}, and braneworld scenario
\cite{Y.Ling} were derived with the help of holographic principle and
the equipartition rule of energy. The Newtonian gravity in loop
quantum gravity was derived in Ref. \cite{Smolin}. In Ref.
\cite{T.Wang} the coulomb force was regarded as an entropic force. In
Ref. \cite{M.Li}, it was shown that the holographic dark energy can
be derived from the entropic force formula. It was pointed out in
\cite{Makela} that Verlinde's entropic force is actually the
consequence of a specific microscopic model of spacetime. The similar
ideas were also applied to the construction of holographic actions
from black hole entropy \cite{Caravellila}. While Ref.
\cite{J.-W.Lee} showed that gravity has a quantum informational
origin.
On the other hand, a modified entropic force in the Debye's model was gave in
\cite{Gao}. In Ref. \cite{L.Zhao}, some of the problems of Verlinde's proposal
on the thermodynamical origin of the principle of relativity was presented, and
the thermodynamic origin of the principle of relativity was explained by hidden
symmetries of thermodynamics. Other applications can be seen in
\cite{Piazza,Culetu,Wang2,WuXN1002.1275}.
According to Verlinde's idea, the holographic screens locate at equipotential
surfaces, where the potential is defined by a timelike Killing vector. Then the
local temperature on a screen can be defined by the acceleration of a particle
that is located very close to the screen. The energy on the screen is
calculated by the holographic principle and the equipartition rule of energy
$E=\frac{1}{2}{\int}TdN$ with $dN$ the bit density of information on the
screen. In this paper, we apply these formulas to investigate the temperature
and energy on holographic screens for 4-dimensional static spherically
symmetric and the Kerr black hole, and look for experiment methods for testing
the entropic force.
The paper is organized as follows. In Sec. \ref{Sec2}, we review Verlinde's
idea about the temperature and the energy from an entropic force in general
relativity. Then, with the idea of entropic force together with the
equipartition law of energy, we calculate the temperature and the energy for
4-dimensional static spherically symmetric black holes and the Kerr black hole
in Sec. \ref{Sec3} and Sec. \ref{Sec4}, respectively. Finally, we give a brief
discussion and present that the entropic force could be tested by experiments.
\section{ The temperature and the energy from an entropic force}
\label{Sec2}
We first review the idea of Verlinde about the temperature and the
energy from an entropic force in general relativity. Consider a
static background with a global time like Killing vector $\xi^\mu$.
One can relate the choice of this Killing vector field with the
temperature and the energy.
In order to define the temperature, we first need to introduce the
potential $\phi$ via the timelike Killing vector $\xi^\mu$:
\begin{eqnarray}
\phi ={1\over 2} \log(-\xi^\mu\xi_\mu),\label{phi}
\end{eqnarray}
where $\xi_\mu$ satisfies the Killing equation
\begin{eqnarray}
\nabla_\mu\xi_\nu+\nabla_\nu\xi_\mu=0.
\end{eqnarray}
The redshift factor is denoted by $e^\phi$ and it relates the local
time coordinate to that at a reference point with $\phi=0$, which we
will take to be at infinity. The potential is used to define a
foliation of space, and the holographic screens are put at surfaces
of constant redshift. So the entire screen has the same time
coordinate.
The four velocity $u^\mu$ and the acceleration
$a^\mu\!\equiv\!u^\nu\nabla_\nu u^\mu$ of a particle that is located
very close to the screen can be expressed in terms of the Killing
vector $\xi^\mu$ as
\begin{eqnarray}
u^\mu &=& e^{-\phi}\xi^\mu, \\
a^\mu &=& e^{-2\phi}\xi^\nu\nabla_\nu\xi^\mu
=-\nabla^{\mu}\phi. \label{amu}
\end{eqnarray}
Note that because the acceleration is the gradient of the potential, it is
perpendicular to screen $\cal S$. So we can turn it into a scalar quantity by
contracting it with a unit outward pointing vector $n^\mu$ normal to the screen
$\cal S$ and to $\xi^\mu$. With the normal vector
$n^\mu=\nabla^\mu\phi/\sqrt{\nabla^\nu\phi\nabla_\nu\phi}$, the local
temperature $T$ on the screen is defined by
\begin{eqnarray}
T=-{\hbar \over 2\pi} e^{\phi} n^\mu a_\mu
={\hbar \over 2\pi} e^{\phi} n^\mu\nabla_\mu\phi
={\hbar \over 2\pi} e^{\phi} \sqrt{\nabla^\mu\phi\nabla_\mu\phi},
\label{T}
\end{eqnarray}
where a redshift factor $e^\phi$ is inserted because the
temperature $T$ is measured with respect to the reference point at
infinity. We will call the temperature defined in (\ref{T}) as
Unruh-Verlinde temperature.
Assuming that the change of entropy at the screen is $2\pi$ for a
displacement by one Compton wavelength normal to the screen, one has
\begin{eqnarray}
\nabla_\mu S = -2\pi {m\over \hbar} n_\mu.
\end{eqnarray}
Now the entropic force, which is required to keep a particle with mass $m$ at
fixed position near the screen, is turned out to be
\begin{eqnarray}
\label{entropicforce}
F_\mu =T\nabla_\mu S = -m e^\phi\nabla_\mu\phi,\label{Fmu}
\end{eqnarray}
where $-\nabla_\mu\phi$ is the relativistic analogue of Newton's acceleration,
and the additional factor $e^\phi$ is due to the redshift, which is a
consequence of the entropy gradient.
We now consider a holographic screen on a closed surface of
constant redshift $\phi$. The number of bit $N$ of the
screen is assumed to proportional to the area of the screen
and is given by \cite{Padmanabhan2010a}
\begin{eqnarray}
N= {A\over \hbar}.
\end{eqnarray}
Then, by assuming that each bit on the holographic screen contributes
an energy $T/2$ to the system, and by using the equipartition law of
energy, we get
\begin{eqnarray} \label{Energy_a}
E={1\over 2} \int_{\cal S} T dN
\end{eqnarray}
with $dN$ the bit density on the screen. Inserting the expressions of
$T$ and $N$ into (\ref{Energy_a}) results in
\begin{eqnarray} \label{Energy}
E ={1\over 4\pi}\int_{\cal S} e^{\phi} \nabla \phi ~ dA.\label{E}
\end{eqnarray}
In the following, we will calculate the Unruh-Verlinde temperature
and the energy on holographic screens for 4-dimensional black
holes with static spherically symmetric metric and stationary
axisymmetric metric, respectively.
\section{Unruh-Verlinde Temperature and energy for static spherically symmetric black holes}
\label{Sec3}
We first consider the case of 4-dimensional general static
spherically symmetric black holes. In the general case, the metric
can be taken as the form
\begin{equation}
ds^2 = -f(r) dt^2 + \frac{dr^2}{f(r)} + r^2 d\Omega_2^2, \label{general1}
\end{equation}
where $d\Omega_2^2 =d\theta^2+\sin^{2}{\theta}d\varphi^2$ is the metric on a
unit 2-sphere. We assume
\begin{equation}
\lim_{r \rightarrow \infty} f(r) = 1 \label{assume_f(r)}
\end{equation}
for ensuring that the metric is asymptotically flat. The event
horizon radius $r_{EH}$ is usually determined by the largest
solution of $f(r)=0$.
By using the Killing equation
\begin{equation}
\partial_\mu\xi_\nu +\partial_\nu\xi_\mu
- 2 \Gamma^{\lambda}_{\mu\nu} \xi_{\lambda}=0,\label{KillingEq}
\end{equation}
and the static spherically symmetric properties of the metric
(\ref{general1})
\begin{equation}
\partial_t \xi_\mu = \partial_\varphi \xi_\mu=0,
\end{equation}
as well as the condition $\xi_\mu\xi^\mu=-1$ at infinity, the
timelike Killing vector of the general black hole is solved as
\begin{eqnarray}
\xi_\mu=\left(-f(r),0,0,0\right),
\end{eqnarray}
which is zero at the event horizon.
According to (\ref{phi}), (\ref{amu}) and (\ref{T}), the
potential, the acceleration and the Unruh-Verlinde temperature on
the holographic screen put at the spherical surface with radius
$r$ are calculated as \footnote{We are grateful to R. A. Konoplya
for pointing out in Ref. \cite{Konoplya} the typo in the numerical
factor of the expression of $a^{r}$ in the first version of our
manuscript. }
\begin{eqnarray}
\phi &=&\frac{1}{2}\ln f(r), \label{phi_spherically}\\
a^\mu &=&\left(0,-\frac{1}{2}f'(r),0,0\right), \label{a_spherically}\\
T &=& \frac{{\hbar}}{4\pi}|f'(r)|. \label{T_spherically}
\end{eqnarray}
The energy on the screen is
\begin{eqnarray}
E = \frac{1}{2}|f'(r)|r^2. \label{E_spherically}
\end{eqnarray}
Next, we focus on the Schwarzschild black hole and the RN black
hole.
\subsection{The Schwarzschild black hole}
For the Schwarzschild black hole, the metric has the form of
(\ref{general1}) with
\begin{equation}
f(r)=1-\frac{2M}{r}.\label{f(r)_Schwarz}
\end{equation}
The event horizon of the balk hole is located at $r_{EH}=2M$, and
the Hawking temperature $T_H$ on the event horizon is given by
\begin{eqnarray}
T_H=\frac{{\hbar}}{4\pi r_{EH}}. \label{Schwarzschild_TH}
\end{eqnarray}
The acceleration and the Unruh-Verlinde temperature on the screen are
calculated as
\begin{eqnarray}
a^\mu &=&\left(0,\frac{M}{r^2},0,0\right), \label{a_Schwarzschild}\\
T &=& \frac{{\hbar}M}{2\pi r^2}. \label{T_Schwarzschild}
\end{eqnarray}
Note that the Unruh-Verlinde temperature on the event horizon
$r=r_{EH}$ is just the Hawking temperature $T_H$:
\begin{eqnarray}
T|_{r=r_{EH}}=\frac{\hbar}{8\pi M}=T_H. \label{RelationWithTh_Sch}
\end{eqnarray}
Then, with the expressions (\ref{E_spherically}) and (\ref{f(r)_Schwarz}), we
obtain the energy on the screen:
\begin{eqnarray}
E = M. \label{E_Schwarzschild}
\end{eqnarray}
It is interesting to note that this energy is independent of the
radius of the screen. In fact this is a consequence of the Gauss'
law.
\subsection{The RN black hole}
As an important example of spherically symmetric black holes, the
RN black hole has the metric form of (\ref{general1}) with
\begin{equation}
f(r)=\left(1-\frac{2M}{r}+\frac{Q^2}{r^2}\right). \label{f(r)_RN}
\end{equation}
There are two horizons, the event horizon with radius $r_+$ and the Cauchy
horizon with radius $r_{-}$, where
\begin{equation}\label{horizonRN}
r_{\pm}= M \pm \sqrt{M^2 -Q^2}.
\end{equation}
The extremal RN black hole corresponds to the case $r_{+} = r_{-}$ or,
equivalently, $M=Q$. The Hawking temperatures $T_{H\pm}$ on the horizons are
\begin{eqnarray}
T_{H\pm}=\frac{{\hbar}}{4\pi}\frac{(r_+ -r_-)}{r_{\pm}^2}
=\frac{{\hbar}}{2\pi}\frac{\sqrt{M^2 -Q^2}}{r_{\pm}^2}.
\end{eqnarray}
The potential, the acceleration and the Unruh-Verlinde temperature
on the screen read
\begin{eqnarray}
\phi &=&\frac{1}{2}\ln\left(1-\frac{2M}{r}+\frac{Q^2}{r^2}\right),
\label{phi_RN}\\
a^\mu &=&\left(0,\frac{Q^2-Mr}{r^3},0,0\right). \label{a_RN} \\
T &=& \frac{{\hbar}}{2\pi}
\frac{|Mr-Q^2|}{r^3}. \label{T_RN}
\end{eqnarray}
On the horizons $r=r_{\pm}$, we have
\begin{eqnarray}
T|_{r=r_{\pm}}=\frac{{\hbar}}{2\pi }
\frac{\sqrt{M^2 -Q^2}}{r_{\pm}^2},
\end{eqnarray}
which shows that the Unruh-Verlinde temperatures on both the horizons equal to
the Hawking temperatures $T_{H\pm}$. The solution of $\partial_r T=0$ is
$r=3q^2/(2M)\equiv r_0$. So for $r_0<r_+$, i.e., $Q^2<8M^2/9$, the maximum of
$T$ in the range $r\geq r_+$ is at $r=r_+$. For $8M^2/9<Q^2<M^2$, the maximum
of $T$ in the range $r\geq r_+$ is at $r=r_0(>r_+)$ and is given by
\begin{eqnarray}
T_{max}=T|_{r_0}=\frac{{\hbar}}{2\pi }
\frac{2M^3}{27Q^4}.
\end{eqnarray}
The energy on the screen is
\begin{eqnarray}
E = \left|M-\frac{Q^2}{r}\right|, \label{E_RN}
\end{eqnarray}
which is the well-known Komar energy of the RN black hole. Now this energy is
dependent on the radius of the screen, which is also a consequence of the
Gauss' laws of gravity and electrostatics. For the screen at infinity, the
energy is reduced to that of the Schwarzschild black hole: $E=M$. It is
interesting to note that the energies on the horizons $r_{\pm}$ are the same:
\begin{eqnarray}
E|_{r=r_{\pm}} = \sqrt{M^2-Q^2}. \label{E_RN_h}
\end{eqnarray}
For the extremal RN black hole with $r_+=r_-$, the corresponding
Unruh-Verlinde temperature and energy are
\begin{eqnarray}
T &=& \frac{{\hbar}}{2\pi}
\frac{M|r-M|}{r^3},\\
E &=& M\left|1-\frac{M}{r}\right|. \label{E_RN2}
\end{eqnarray}
They vanish at the horizon $r_+=M$. This result is in agreement with the
statements that the extremal black hole has a unique internal state
\cite{Hawking1995} and its temperature is zero due to the vanishing of surface
gravity on horizon.
\section{Unruh-Verlinde Temperature and energy for the Kerr black hole}
\label{Sec4}
In this section, we extend this work to the Kerr black hole. The Kerr solution
describes both the stationary axisymmetric asymptotically flat gravitational
field outside a massive rotating body and a rotating black hole with mass and
angular momentum. The Kerr black hole can also be viewed as the final state of
a collapsing star, uniquely determined by its mass and rate of rotation.
Moreover, its thermodynamical behavior is very different from the Schwarzschild
black hole and the RN black hole, because of its much more complicated causal
structure. Hence its study is of great interest in understanding physical
properties of astrophysical objects, as well as in checking any conjecture
about thermodynamical properties of black holes.
In terms of Boyer--Lindquist coordinates, the Euclidean Kerr metric
reads
\begin{eqnarray}
ds^{2} &=& -\left(1-\frac{2Mr}{\varrho^{2}}\right)dt^2
+ \frac{\varrho^{2}}{\Delta}dr^{2}
+\varrho^{2}d\theta^{2} \nonumber\\
&& +\left((a^2+r^2)\sin^{2}\theta
+\frac{2Mra^2\sin^{4}\theta}{\varrho^{2}}
\right)d\varphi^2 -\frac{4Mra\sin^{2}\theta}{\varrho^{2}}dtd\varphi,
\label{kermet}
\end{eqnarray}
where $\Delta$ is the Kerr horizon function
\begin{eqnarray}
\Delta = r^{2}-2Mr +a^{2},
\end{eqnarray}
and
\begin{eqnarray}
\varrho = r^{2} + a^{2}\cos^{2}\theta.
\end{eqnarray}
Here $a$ is the angular momentum for unit mass as measured from
the infinity; it vanishes in the Schwarzschild limit. The
nonextremal Kerr black hole has the event horizon $r_{+}$ and the
Cauchy horizon $r_{-}$ at
\begin{equation}
r_{\pm}=M \pm \sqrt{M^2-a^2}.
\end{equation}
The extreme case corresponds to $r_{+}=r_{-}$ or $M=a$. The Hawking
temperatures on the horizons are
\begin{eqnarray}
T_{H\pm}=\frac{{\hbar}}{4\pi}\frac{(r_+ -r_-)}{(r_{\pm}^2+a^2)}
=\frac{{\hbar}}{4\pi}\frac{\sqrt{M^2 -a^2}}{Mr_{\pm}}.
\end{eqnarray}
The solution for the timelike Killing vector $\xi_\mu$ is read as
\cite{WuXN1002.1275}
\begin{eqnarray}
\xi_\mu=\left(-1+\frac{2Mr(\Omega-2a^2 Mr\sin^2\theta)}
{\Omega\varrho^{2}},0,0,0\right),
\label{KillingVectorKerr}
\end{eqnarray}
where $\Omega=(r^2+a^2)^2-a^2{\Delta}\sin^2\theta$. The corresponding
potential is
\begin{eqnarray}
\phi=-\frac{1}{2}\log\left({1+\frac{2Mr(r^2+a^2)}{\Delta\varrho^2}}\right),
\label{PotentialKerr}
\end{eqnarray}
which shows that equipotential surfaces are dependent with two parameters $r$
and $\theta$, hence the holographic screens are not spherically symmetric but
axisymmetric. However, when $r\gg r_+$ or $r\rightarrow r_{\pm}$, the
holographic screens are approximate spherically symmetric, which can be seen
from the contour plots of the potential $\phi$ in the $y$-$z$ plane showed in
Figs. \ref{fig:ContourPhiA} and \ref{fig:ContourPhiB}. It can be seen that,
when $a\rightarrow M$, the effect of the angular momentum are remarkable near
the horizons. When far away from the event horizon, the equipotential line can
be approximated as a spherical surface.
\begin{figure}[h]
\begin{center}
\subfigure[$r\geq r_+=16$]{\label{fig:ContourPhiAOut}
\includegraphics[width=6.5cm]{Phi_Kerr_10_8_OutNodata.eps}}
\subfigure[$r\leq r_-=4$]{\label{fig:ContourPhiAIn}
\includegraphics[width=6.5cm]{Phi_Kerr_10_8_InNodata.eps}}
\end{center}\vskip -4mm
\caption{ A contour plot of the potential $\phi$
in the $y$-$z$ plane for the Kerr
black hole. The parameters are set to $M=10$ and $a=8$.}
\label{fig:ContourPhiA}
\end{figure}
\begin{figure}[h]
\begin{center}
\subfigure[$r\geq r_+=11.4$]{\label{fig:ContourPhiBOut}
\includegraphics[width=6.5cm]{Phi_Kerr_10_9.9_OutNodata.eps}}
\subfigure[$r\leq r_-=8.6$]{\label{fig:ContourPhiAIn}
\includegraphics[width=6.5cm]{Phi_Kerr_10_9.9_InNodata.eps}}
\end{center}\vskip -4mm
\caption{ A contour plot of the potential $\phi$
in the $y$-$z$ plane for the Kerr
black hole with $a\rightarrow M$. The parameters are set to
$M=10$ and $a=9.9$.}\label{fig:ContourPhiB}
\end{figure}
The non-zero components of the acceleration are
\begin{eqnarray}
a^r &=& \frac{2Mr(a^2+r^2)
\left(M-r-\frac{\Delta r}{\varrho^2}\right)
+M\Delta(a^2+3r^2)}
{\varrho^2[\Delta\varrho^2+2Mr(a^2+r^2)]}, \nonumber \\
a^\theta &=& \frac{Mra^2(a^2+r^2)\sin(2\theta)}
{\varrho^4[2Mr(a^2+r^2)+\Delta\varrho^2]}.
\label{a_Kerr}
\end{eqnarray}
The Unruh-Verlinde temperature is given by
\begin{eqnarray}
T&=&\frac{h}{2\pi } \left\{
\frac{M^2\big[\left(a^2+3 r^2\right) \Delta \varrho ^2+2 r
\left(a^2+r^2\right)
\left(M \varrho ^2-r \left(\Delta
+\varrho^2\right)\right)\big]^2 }
{\varrho ^4 \left[2 M r \left(a^2+r^2\right)
+\Delta \varrho ^2\right]^3} \right. \nonumber \\
&& \quad\quad~ + \left. \frac{\Delta
\left[a^2 M r (a^2+r^2)
\texttt{sin}(2\theta)
\right]^2 }
{\varrho ^4 \left[2 M r \left(a^2+r^2\right)
+\Delta \varrho ^2\right]^3} \right\}
^{\frac{1}{2}} .
\label{T_Kerr}
\end{eqnarray}
On the horizons $r=r_{\pm}$ ($\Delta=0$), the Unruh-Verlinde temperatures
become
\begin{eqnarray}
T|_{r=r_{\pm}}&=&\left. \frac{\hbar}{2\pi }\sqrt{
\frac{(M -r)^2 }{2 M r (a^2+r^2)} }~ \right|_{r=r_{\pm}}\nonumber \\
&=& \frac{{\hbar}}{4\pi}
\frac{\sqrt{M^2-a^2}} {Mr_{\pm}}
=T_{H\pm}.
\label{T_Kerr_horizons}
\end{eqnarray}
So, the Unruh-Verlinde temperatures on both horizons are equal to the Hawking
temperatures. For vanishing $a$, the above result is reduced to
$T|_{r=r_+}=\frac{{\hbar}}{4\pi r_+}$, which is just the Hawking temperature
(\ref{T_Schwarzschild}) on the horizon of the Schwarzschild black hole. Contour
plots of the temperature $T$ in the $y$-$z$ plane are shown in Figs.
\ref{fig:ContourTA} and \ref{fig:ContourTB}.
\begin{figure}[h]
\begin{center}
\subfigure[$r\geq r_+=16$]{\label{fig:ContourTAout}
\includegraphics[width=6.5cm]{T_Kerr_10_8_OutNodata.eps}}
\subfigure[$r\leq r_-=4$]{\label{fig:ContourTAin}
\includegraphics[width=6.5cm]{T_Kerr_10_8_InNodata.eps}}
\end{center}\vskip -4mm
\caption{ A contour plot of the temperature $T$
in the $y$-$z$ plane for the Kerr black
hole. The parameters are set to $M=10$ and
$a=8$.}\label{fig:ContourTA}
\end{figure}
\begin{figure}[h]
\begin{center}
\subfigure[$r\geq r_+=11.4$]{\label{fig:ContourTBout}
\includegraphics[width=6.5cm]{T_Kerr_10_9.9_OutNodata.eps}}
\subfigure[$r\leq r_-=8.6$]{\label{fig:ContourTBin}
\includegraphics[width=6.5cm]{T_Kerr_10_9.9_InNodata.eps}}
\end{center}\vskip -4mm
\caption{ A contour plot of the temperature $T$
in the $y$-$z$ plane for the Kerr black hole
with $a\rightarrow M$. The parameters are set to $M=10$ and
$a=9.9$.}\label{fig:ContourTB}
\end{figure}
Now, by using the equipartition law of energy and the holographic principle
$E={1\over 2} \int_{\cal S} T dN=\frac{1}{2\hbar} \int_{\cal S} T dA$,
we get the energies on the horizons
\begin{eqnarray}
E|_{r=r_{\pm}}
&=& \frac{\sqrt{M^2-a^2}}{2M r_{\pm}}\left( r_{\pm}^2+a^2 \right) \nonumber \\
&=& \sqrt{M^2-a^2}, \label{E_Kerr_h}
\end{eqnarray}
which is the reduced mass $M_0$ of the Kerr black hole. When far away from the
event horizon, or equivalently $r\gg r_+$, the screen is approximated as a
sphere and the energy can be expressed as
\begin{eqnarray}
E
&\approx& M\left(1-4\frac{a^2M}{r^3}\right). \label{E_Kerr_Approx}
\end{eqnarray}
For $r\rightarrow\infty$, the result is $E=M$, which is independent of the
angular of the Kerr black hole. For $a=0$, the expression (\ref{E_Kerr_Approx})
reduce to $E=M$, which is just the energy expression (\ref{E_Schwarzschild})
for the Schwarzschild black hole.
\section{Discussions}\label{SecConclusions}
In this paper, with the holographic principle and the equipartition law of
energy, we investigate the Unruh-Verlinde temperature and energy on holographic
screens for several 4-dimensional black holes with static spherically symmetric
metric and stationary axisymmetric metric.
On the event horizon of a static spherically symmetric black hole, the Hawking
temperature $T_H$ is related to the surface gravity
$\kappa=\frac{1}{2}f'(r_{EH})$ by the relation
$T_H=\frac{\hbar\kappa}{2\pi}=\frac{{\hbar}}{4\pi}f'(r_{EH})$, while the
Unruh-Verlinde temperatures $T$ is related to the acceleration $a^\mu
=\left(0,-\frac{f'(r)}{2},0,0\right)$ by $T=-{\hbar \over 2\pi} e^{\phi} n^\mu
a_\mu=\frac{{\hbar}}{4\pi}f'(r_{EH})$, which shows that the Unruh-Verlinde
temperatures $T$ for a static spherically symmetric black hole is identical to
the Hawking temperature. Hence, outside the horizons, the Unruh-Verlinde
temperature $T$ can be considered as a generalized Hawking temperature on the
holographic screen. For the Kerr black case, we also consider $T$ as the
generalized Hawking temperature. On the screen located at the infinity, the
surface gravity or the acceleration of a particle (which is free there)
vanishes, hence the temperature is zero.
The entropic force (\ref{Fmu}) for a static spherically symmetric black hole
can be calculated as $F_{\mu}=(0,-\frac{mf'(r)}{2\sqrt{f(r)}},0,0)$. The
magnitude of the force is
$F=\sqrt{g^{\mu\nu}F_{\mu}F_{\nu}}=\frac{1}{2}mf'(r)$. For the Schwarzschild
black hole, the entropic force $F$ is just the Newton force
\begin{equation}
F=\frac{GmM}{r^2}, \label{F_Schwarzschild}
\end{equation}
here we recover the Newton's gravitational constant $G$. While, for the RN
case, the entropic force is turned out to be
\begin{equation}
F=\frac{GmM}{r^2}\left(1-\frac{Q^2}{Mr}\right), \label{F_RN}
\end{equation}
which shows that there is a corrected term caused by the charge of the black
hole. This force is related to the pure gravitational effect. The gravity or
the geometry near the event horizon of the black hole is affected by the energy
of the electric field. When far away form the horizon, where the electric field
becomes weak, the entropic force will recover to the Newton one. Note that,
since $r\geq r_+>M>Q$, we have $Q^2/(Mr)<1$ and $F>0$. Similarly, the angular
momentum of the Kerr black hole will also affect the structure of the spacetime
near the event horizon. The entropic force of the Kerr black hole is read as
\begin{eqnarray}
F=\frac{Gm\sqrt{M^2-a^2}}{r^2}\label{F_Kerr_1}
\end{eqnarray}
and
\begin{eqnarray}
F\approx\frac{GmM}{r^2}\left(1-4\frac{a^2M}{r^3}\right)
\label{F_Kerr_2}
\end{eqnarray}
at $r=r_+$ and $r\gg r_+$, respectively. It also reduces at large $r$ to the
Newton force because the effect of the ``field" caused by the angular momentum
of the black hole becomes weaker with the increase of $r$.
The energy got from the holographic principle is indeed the Komar energy on the
screen. With the same explanation above, we also can understand that the energy
$E$ on the screen will trend to the ADM mass $M$ when $r\rightarrow\infty$ for
anyone of the black holes considered in the paper. This is because the ``field"
produced by the charge or the angular momentum is strong on the screen near the
event horizon and weak when far away from the black hole. On the screen at
infinity, the effect of them vanishes and the energy is naturally equal to the
black hole mass $M$ according to the gravitational Gauss' law. We see clearly
from (\ref{E_RN_h}) and (\ref{E_Kerr_h}) that when applying the entropic force
idea to the black hole horizon, it is the reduced mass $M_0$ that takes the
place of the black hole mass $M$. For the Schwarzschild black hole, there is no
other parameter beside $M$, so the reduced mass $M_0$ is just the black hole
mass $M$. For the RN black hole, the reduced mass is read from (\ref{E_RN}) as
$M_0=M-\frac{Q^2}{r}$, which is $M_0=\sqrt{M^2-Q^2}$ on the event horizon. For
the Kerr case, the reduced mass is $M_0=\sqrt{M^2-a^2}$ on the event horizon
and is $M_0 {\approx} M \left(1-4\frac{a^2M}{r^3}\right)$ at $r\gg r_+$.
Therefore, with the black hole mass $M$ replaced by the reduced mass $M_0$, we
can rewrite the entropic force of a black hole as
\begin{equation}
F=\frac{GmM_0}{r^2}, \label{F_Entropic}
\end{equation}
which has the same form as the Newton's law of gravity. This formula shows that
the force between two neutral particles is exactly the Newton's force of
gravity, while the force between a neutral particle and a charged particle
would departure from the Newton's law of gravity. Hence, the idea of entropic
force could be tested by experiments according to Eq. (\ref{F_Entropic}).
Explicitly, the force between a neutral particle and a charged particle is $F=\frac{Gm}{r^2}\left(M-\frac{Q^2}{r}\right)$ according entropic force idea. Therefore, by comparing the high accurate experiment results of the force between a neutral particle and a charged particle with the theoretical predictions of the entropic force, one could check whether the entropic idea is right or wrong.
\section*{Acknowledgment}
This work was supported by the Program for New Century Excellent Talents in
University, the National Natural Science Foundation of China (No. 10705013),
the Key Project of Chinese Ministry
of Education (No. 109153), and the Fundamental Research Funds for the Central
Universities (No. lzujbky-2009-54, No. lzujbky-2009-122 and No.
lzujbky-2009-163).
\section*{References}
|
\section{Introduction }
In recent years, primary processes in photosynthesis have received
a renewed interest from a broader physical community thanks to experimental
observation of coherent energy transfer in some photosynthetic systems.
The ground breaking coherent two-dimensional electronic spectroscopy
(2D-ES) experiment of Engel \emph{et al.} \cite{Engel2007a} has led
to new appreciations of the role that may be played by coherent dynamics
in excitation energy transfer (EET), and of the quantum mechanical
nature of photosynthetic systems in general \cite{Scholes2010a}.
Special theoretical effort has been made to understand the role of
noise \cite{Darius2004,Plenio2008,Castro2008,Rebentrost2009,Mohseni2009}
in the dynamics of excitation energy transfer, and the role of coherence
\cite{Ishizaki1,Ishizaki2,Ishizaki-PNAS,Cheng-Review,Olsina2010a}
in excitonicaly coupled systems. On the experimental front, the method
of coherent 2D-ES \cite{Jonas2003a,Cho2008a} has established itself
as a tool opening new window into the details of energy transfer dynamics
in photosynthetic \cite{Brixner2005a,Cho2005a,Calhoun2009,Cohen2009,Ginsberg2009,Sanda2009},
and other molecular systems \cite{Milota2009,Darius2004,Collini2009}.
Coherent effects have been now reported in different molecular systems,
often biologically relevant \cite{Lee2007a,Collini2009} - a generality
that asks for a search of the possible evolutionary advantage underlying
their abundance in photosynthetic pigment-protein complexes.
The principle pigment molecules responsible for the primary processes
of photosynthesis are chlorophylls (Chls) and bacteriochlorophylls
(BChls) \cite{Blankenship,Chls}. They are involved in accumulation
of light energy via the excitation energy transfer to specific pigment-protein
complexes - reaction centers. Spectral variability of photosynthetic
light-harvesting pigment-protein complexes arises either from excitonic
interactions between pigment molecules or from their interactions
with protein surrounding. Both these interactions are the main factors
determining the excitation dynamics in light-harvesting \cite{AVG}.
Excitonic aggregates are subject to interaction with two types of
environments, and they provide means of transferring energy from one
environment to another. First of these environments, the radiation,
is under natural conditions at much higher temperature than the second
environment, the protein scaffold and indeed the photosynthetic chemical
machinery as a whole. The excess of photons on suitable wavelength
in the radiational environment is used to excite spatially extended
antenna systems that concentrate excitation energy to the reaction
center, which in turn drives charge transfer processes across cellular
membranes to create the transmembrane potential and the pH gradient
\cite{Blankenship}.
Non-equilibrium processes occurring in photosynthetic systems during
light harvesting are conveniently described by reduced density matrix
(RDM) theory \cite{AVG,MayKuehnBook,Bruggemann07} which has an advantage
of being applicable to disordered statistical ensembles that the experiments
often deal with. However, with recent 2D experiments that enable us
to distinguish the homogeneous and inhomogeneous spectral broadening,
and with the progress in single molecular spectroscopy \cite{SingleBook}
we can gain insight into the time evolution characteristic to single
molecules interacting with their environment \cite{Valkunas2007,Janusonis2008}.
This fact enables us to return to the wavefunction formalism and to
look at light harvesting from the point of view which takes the superposition
principle of quantum mechanics seriously. It has been show that such
an approach yields many interesting insights into the emergence of
the classical properties of molecular system from their underlying
quantum mechanical nature \cite{SchlosshauerBook,DecoherenceBook}.
As the light-harvesting processes seem to operate on the interface
between classical and quantum worlds it seems appropriate to look
at them from the point of view of the decoherence program of Zeh,
Zurek and others \cite{Zeh70,Zurek82}.
The process of light harvesting could then be describes as follows.
First, the system is in an {}``equilibrium'' initial state $|\Psi_{0}\rangle$
characterized by the excitonic ground state $|g\rangle$, the state
of protein (phonon) environment $|\Phi_{P}\rangle$ corresponding
to this electronic ground state and some state of light $|\Xi_{0}\rangle$,
i.e. \begin{equation}
|\Psi_{0}\rangle=|g\rangle|\Phi_{B}\rangle|\Xi_{0}\rangle.\label{eq:state_ini}\end{equation}
The light-harvesting occurs when the state of light is such that the
time evolution of the system leads to population of higher excited
states $|e_{n}\rangle$ of photosynthetic antenna. These states are
formed from excited states of Chls and other chromophores, such as
carotenoids \cite{AVG}. We denote these combined excited states as
excitons. In the first approximation, photosynthetic antenna remains
in the excited state until the excitation energy is transferred to
the reaction center. This happens much faster than competing process
of spontaneous emission which can therefore be neglected in our discussion.
When the interaction of the antenna with light is switched on, the
change occurring in the ground state portion of the total state vector
after the passage of time $\Delta t$ is\begin{equation}
|\Psi_{0}\rangle\rightarrow\alpha_{\Delta t}|g\rangle|\Phi_{B}\rangle|\Xi_{0}\rangle+\sum_{n}\beta_{\Delta t}^{(n)}|e_{n}\rangle|\Phi_{B}\rangle|\Xi^{\prime}\rangle\label{eq:state_exc}\end{equation}
The subsequent time evolution of the excited state portion of the
state vector is independent of the ground state part, and we can thus
look at it separately. Because we neglect spontaneous emission, any
excitation to higher excited state, as well as transitions between
exciton states due to the light, the state vector $|\Xi^{\prime}\rangle$
remains approximately unentangled with excitons and the protein bath
for the rest of the energy transfer process. It can therefore be omitted.
The initial state for the energy transfer process thus reads\begin{equation}
|\Psi_{e}(t_{0})\rangle=\sum_{n}\beta^{(n)}(t_{0})|e_{n}\rangle|\Phi_{B}\rangle,\label{eq:ini_in_ex}\end{equation}
where we omitted the lower index $\Delta t$. If the basis of the
states $|e_{n}\rangle$ is chosen so that the molecular Hamiltonian
is diagonal, the energy transfer occurs only due to interaction of
excitons with their surrounding environment. This interaction leads
to an entanglement of excitons and the environment \begin{equation}
|\Psi_{e}(t)\rangle=\sum_{n}\beta^{(n)}(t)|e_{n}\rangle|\Phi_{B}^{(n)}(t)\rangle.\label{eq:entanglement}\end{equation}
After a sufficiently long time the environment state vectors corresponding
to different electronic state diverge maximally and the reduced density
matrix becomes diagonal in some basis, i.e. \[
\rho(t)=tr_{B}\{|\Psi_{e}(t)\rangle\langle\Psi_{e}(t)|\}\]
\begin{equation}
=\sum_{mn}\beta^{(n)}(t)(\beta^{(m)}(t))^{*}\langle\Phi_{B}^{(m)}(t)|\Phi_{B}^{(n)}(t)\rangle.\label{eq:dens_mat_from_ex}\end{equation}
Often, to a good approximation, such preferred basis is the one in
which the electronic Hamiltonian is diagonal, the so-called excitonic
basis. However, notable corrections to this rule are predicted even
for weak system-bath coupling \cite{Tannor,Olsina2010a}.
The final state of the energy transfer is the one in which just reaction
centers are populated\begin{equation}
|\Psi_{e}(t)\rangle=\sum_{k}\beta^{(RC_{k})}(t)|e_{RC_{k}}\rangle|\Phi_{B}^{(RC_{k})}(t)\rangle.\label{eq:in_RC}\end{equation}
The last step of the energy transfer, from the antenna to the reaction
center is often slower than typical transfer times between antenna
complexes, and so the final state is well localized on the reaction
center, and coherences between individual reaction centers are unlikely
to survive.
It is clear from the above discussion, that decoherence during the
energy transfer in the antenna is determined by the evolution of the
environmental degrees of freedom (DOF). The decoherence from the rest
of the system might be required for the localization of the energy
in the reaction center, but there is no obvious reason for fast decoherence
during the initial steps of energy transfer in the antenna, apart
from the fact that a bath formed by a completely random disordered
environment would lead to just such fast decoherence. It has been
suggested before that the protein environment might play a more active
role in steering and protecting electronic excitation \cite{Engel2007a,Lee2007a}
and controlling the decoherence might be one of the possible pathways
to more robust EET.
There is however one important caveat in the above scheme. The initial
condition, Eq. (\ref{eq:ini_in_ex}), has been introduced artificially
into Eq. (\ref{eq:state_exc}) as a result of an interaction occurring
during some short time interval $\Delta t$. If the system is continuously
pumped, individual contributions similar to Eq. (\ref{eq:ini_in_ex})
will interfere, possibly disabling any effect of cooperative involvement
of the bath. It is even more important to consider the question what
are the effects of natural sun light \cite{Brumer}, i.e. whether
the coherent scenario outlined above is plausible for the photosynthetic
system \emph{in vivo,} or not. This depends strongly on the nature
of the excitation process, whether it occurs in discrete independent
jumps of the kind described by Eq. (\ref{eq:state_exc}), or continuously
over a long period of uncertainty interval of the photon arrival.
The former view is usually held in support of the relevance of ultrafast
spectroscopic finding for \emph{in vivo} function of the photosynthetic
systems \cite{Cheng-Review}. Below we derive a general formula which
enables us to describe all these regimes by a unified formalism, and
also enables us to place the observables of ultrafast coherent spectroscopy
in perspective with the dynamics under natural conditions. In a somewhat
extended form our result is also applied to another case cited in
support of the utility of coherent dynamics in photosynthetic systems,
a case where a small photosynthetic complex is excited through another,
possibly mesoscopic, antenna \cite{Cheng-Review}.
The paper is organized as follows. Next section introduces a rather
general model of photosynthetic aggregate. In Section \ref{sec:Excitation-by-Coherent}
we discuss the dynamics of a system excited by coherent pulsed light
and the observables of the ultra-fast non-linear spectroscopy. Section
\ref{sec:Excitation-by-Thermal} is concerned with the excitation
of a photosynthetic system by the light from a general source. Implications
of the theory for excitation by thermal and coherent light, as well
as excitation mediated by mesoscopic system are discussed in Section
\ref{sec:Discussion}.
\section{Hamiltonian of a Model Photosynthetic System\label{sec:Hamiltonian-of}}
In this section we briefly review the excitonic model that was very
successfully applied to model the spectroscopic properties of Chl-
and Bchl- based light harvesting chromophore--protein complexes (see
e.g. \cite{Cho2005a}). We assume $N$ monomers with ground states
$|g_{n}\rangle$, excited states $|\tilde{e}_{n}\rangle$, $n=1,\dots,N$,
and with electronic transition energies $\tilde{\varepsilon}_{n}$.
These monomers are interacting with the phonon bath of protein DOF
so that the Hamiltonian of the monomer reads\begin{equation}
H_{n}=(T+V_{g})|g_{n}\rangle\langle g_{n}|+(\tilde{\varepsilon}_{n}+T+V_{e})|\tilde{e}_{n}\rangle\langle\tilde{e}_{n}|.\label{eq:mono-ham}\end{equation}
Here, $T$ is the kinetic energy operator of the bath, and $V_{g}$
and $V_{e}$ are the potential energy operators of the bath when the
system is in the electronic ground- and excited states, respectively.
We set the ground state electronic energy to zero for conveniency.
The Hamiltonian, Eq. (\ref{eq:mono-ham}) can be split into the pure
bath, pure electronic and the interaction terms so that\[
H_{n}=\underbrace{(T+V_{g})\otimes\mathbb{I}_{M}^{n}}_{H_{B}^{n}}+\underbrace{\mathbb{I}_{B}^{n}\otimes(\tilde{\varepsilon}_{n}+\langle V_{e}-V_{g}\rangle_{eq})|\tilde{e}_{n}\rangle\langle\tilde{e}_{n}|}_{H_{M}^{n}}\]
\begin{equation}
+\underbrace{(V_{e}-V_{g}-\langle V_{e}-V_{g}\rangle_{eq})\otimes|\tilde{e}_{n}\rangle\langle\tilde{e}_{n}|}_{H_{M-B}^{n}}.\label{eq:Ham_n_splitted}\end{equation}
Here, $\mathbb{I}_{B}$ is the unity operator on the bath Hilbert
space and $\mathbb{I}_{M}$ is the unity operator on the Hilbert space
of the electronic states. The equlibrium average $\langle V_{e}-V_{g}\rangle_{eq}$
of the potential energy operators was added to the electronic energy
so that the interaction term is zero for the system in equilibrium.
In chromophore--protein complexes many such monomers are coupled by
resonance coupling. The whole complex can be described by means of
collective states including the ground state\begin{equation}
|g\rangle=\prod_{n=1}^{N}\otimes|g_{n}\rangle,\label{eq:ground}\end{equation}
one excitation states\begin{equation}
|\bar{e}_{a}\rangle=\prod_{n=1}^{a-1}\otimes|g_{n}\rangle\otimes|\tilde{e}_{a}\rangle\otimes\prod_{m=a+1}^{N}\otimes|g_{m}\rangle,\label{eq:oneexc}\end{equation}
and states containing higher number of excitations. For the sake of
brevity we now stop writing the symbol of the direct product $\otimes$
and the unity operators $\mathbb{I}_{B}^{n}$ etc. explicitely. The
total Hamiltonian of the complex including resonance interaction is
then defined as\[
H_{B}+H_{M}+H_{M-B}=\sum_{n=1}^{N}H_{B}^{n}+\sum_{n=1}^{n}H_{M}^{n}\]
\begin{equation}
+\sum_{n\neq m}^{N}J_{nm}|\bar{e}_{n}\rangle\langle\bar{e}_{m}|+\sum_{n=1}^{n}H_{M-B}^{n}.\label{eq:Htot}\end{equation}
If the system-bath interaction with bath is weak, the referred basis
into which the electronic system relaxes due to interaction with the
bath is, to a good approximation, the one in which the electronic
part of the Hamiltonian is diagonal. Let us denote these states as
$|e_{n}\rangle$. They are usually termed excitons and they represent
certain linear combination of the collective states $|\bar{e}_{n}\rangle$
where excitations are localized on individual chromophore molecules.
One of the most important characteristics of this model is that it
does not include direct relaxation of the electronic excited states
to the ground states due to electron-phonon coupling. This is well
satisfied by Chls and BChls on the ultrafast time scale of which light
harvesting processes occur.
\section{Excitation by Coherent Pulsed Light and Non-Linear Spectroscopy\label{sec:Excitation-by-Coherent}}
Let us now consider experimental methods which provide information
about time evolution of excited states of photosynthetic systems.
Because of the timescale of EET processes, spectroscopy with ultrashort
time resolution is a necessary tool. The interaction of the pulsed
coherent light with the photosynthetic system is well described in
semi-classical approximation \cite{MukamelBook}. Electric field of
the light is then considered as an external parameter of the system
Hamiltonian. Electronic DOF can be prepared very fast in an excited
state, not affecting, to a good approximation, the bath DOF. Thus,
in an experiment with an ideal time resolution, we would have the
system prepared in the excited state, Eq. (\ref{eq:ini_in_ex}). The
time evolution of the system is governed by the Schr\"{o}dinger equation\[
\frac{\partial}{\partial t}|\Psi_{e}(t)\rangle=-\frac{i}{\hbar}(H_{B}+H_{M}+H_{M-B})|\Psi_{e}(t)\rangle\]
\begin{equation}
+\delta(t)|\Psi_{e}(t_{0})\rangle,\label{eq:with_delta}\end{equation}
with initial condition $|\Psi_{e}(t)\rangle=0$ for $t<t_{0}$. The
last term in Eq. (\ref{eq:with_delta}) describes the ultrafast event
of the molecule--radiation interaction. Formal solution of this equation
reads $|\Psi_{e}(t)\rangle=U_{B}(t)U_{M}(t)U_{M-B}(t)|\psi_{e}(t_{0})\rangle$,
where we defined evolution operators $U_{B}(t)$, $U_{M}(t)$ of the
bath and the molecule, respectively, as \begin{equation}
U_{B}(t)=\Theta(t-t_{0})\exp\{-\frac{i}{\hbar}H_{B}(t-t_{0})\},\label{eq:evol_op0}\end{equation}
\begin{equation}
U_{M}(t)=\Theta(t-t_{0})\exp\{-\frac{i}{\hbar}H_{M}(t-t_{0})\},\label{eq:evol_op}\end{equation}
and the remaining interaction evolution operator as\[
U_{M-B}(t)=\Theta(t-t_{0})\exp\{-\frac{i}{\hbar}\int\limits _{t_{0}}^{t}d\tau U_{B}^{\dagger}(\tau)U_{M}^{\dagger}(\tau)\]
\begin{equation}
\times H_{M-B}U_{M}(\tau)U_{B}(\tau)\},\label{eq:sol_with_delta}\end{equation}
After excitation, the process of energy transfer proceeds according
to the description presented in Introduction and can be experimentally
monitored.
\subsection{Evolution superoperator}
Matrix elements of the RDM of the molecule, which holds the information
about the population probabilities and the amount of coherence between
electronic states are given by expectation value of projectors $|e_{n}\rangle\langle e_{m}|$,
\[
\rho_{nm}(t)=\langle\psi_{e}(t)|e_{m}\rangle\langle e_{n}|\psi_{e}(t)\rangle\]
\[
=tr_{B}\{\langle e_{n}|\psi_{e}(t)\rangle\langle\psi_{e}(t)|e_{m}\rangle\}\]
\[
=\langle e_{n}|tr_{B}\{U_{M}(t)U_{M-B}(t)\underbrace{\sum_{ab}\beta^{(a)}(\beta^{(b)})^{*}|a\rangle\langle b|}_{\rho_{0}}\]
\begin{equation}
\times\underbrace{|\Phi_{B}\rangle\langle\Phi_{B}|}_{W_{eq}}U_{M-B}^{\dagger}(t)U_{M}^{\dagger}(t)\}|e_{m}\rangle\label{eq:P_nmt}\end{equation}
This can be rewritten by defining an evolution superoperator ${\cal U}(t)$
which acts on initial density matrix $\rho_{0}W_{eq}$, i.e.\[
\rho_{nm}(t)=tr_{B}\{\langle e_{n}|W(t)|e_{m}\rangle\}\]
\begin{equation}
=\langle e_{n}|{\cal U}^{(e)}(t)\rho_{0}W_{eq}|e_{m}\rangle,\label{eq:P_nm_sup}\end{equation}
The matrix elements of the superoperator read \[
{\cal U}_{abcd}^{(e)}(t)=\langle a|U_{M}(t)U_{M-B}(t)|c\rangle\dots\]
\begin{equation}
\times\langle d|U_{M-B}^{\dagger}(t)U_{M}^{\dagger}(t)|b\rangle,\label{eq:sup_elements}\end{equation}
where the dots $\dots$ denote where an operator on which ${\cal U}^{(e)}(t)$
acts has to be inserted. The reduced evolution operator $\bar{{\cal U}}^{(e)}(t)$
defined as\begin{equation}
\bar{{\cal U}}^{(e)}(t)=tr_{B}\{{\cal U}^{(e)}(t)\},\label{eq:Ue_red}\end{equation}
contains information about the evolution of the RDM only.
\subsection{Non-linear spectroscopy}
In non-linear spectroscopy, coherent laser light is used to investigate
the dynamics of molecular systems by applying special sequences of
pulses. Some pulses act to induce non-equilibrium dynamics (pump),
and other pulses act to monitor (probe) the evolution after the pump.
One of the most advanced of these methods, coherent 2D-ES \cite{Jonas2003a,ChoBook}
measures the response of a system to three pulses traveling in different
directions $\bm{k}_{1}$, $\bm{k}_{2}$ and $\bm{k}_{3}$. The detection
is arranged in such a way (measuring in the direction $-\bm{k}_{1}+\bm{k}_{2}+\bm{k}_{3}$)
that the signal is predominantly of the third order, with contributions
of one order per pulse \cite{MukamelBook}. Let us denote delays between
the first two pulses by $\tau$ and the delay between the second and
the third pulse by $T$. If the pulses are ideally short, the signal
is composed of two kinds of contribution. First, contribution that
involves population of the excited state corresponding to the density
operator \begin{equation}
W_{\bm{k}_{2},\bm{k}_{1}}^{(e)}(t,\tau)=|\psi_{e}^{(\bm{k}_{2})}(t)\rangle\langle\psi_{e}^{(\bm{k}_{1})}(t+\tau)|,\label{eq:We_1}\end{equation}
and second, contribution that involves evolution in the ground state
\begin{equation}
W_{\bm{k}_{2},\bm{k}_{1}}^{(g)}(t,\tau)=|\Psi_{B}\rangle|g\rangle\langle\psi_{g}^{(\bm{k}_{2},\bm{k}_{1})}(t,t+\tau)|.\label{eq:Wg_1}\end{equation}
Here, we denote the pulses acting on the state vector by their corresponding
wave vector in the upper index, and the excited state or ground state
bands by the lower index $g$ and $e$, respectively. For these statistical
operators we can define evolution superoperators ${\cal U}^{(e)}(t)$,
${\cal U}^{(g)}(t)$, ${\cal U}^{(eg)}(t)$ and $U^{(ge)}(t)$ in
analogy with Eqs. (\ref{eq:P_nm_sup}) and (\ref{eq:sup_elements}),
so that\begin{equation}
W_{\bm{k}_{2},\bm{k}_{1}}^{(e)}(t,\tau)={\cal U}^{(e)}(t){\cal U}^{(ge)}(\tau)|\Psi_{B}\rangle\langle\Psi_{B}|\mu|g\rangle\langle g|\mu,\label{eq:We_2}\end{equation}
and \begin{equation}
W_{\bm{k}_{2},\bm{k}_{1}}^{(g)}(t,\tau)={\cal U}^{(g)}(t){\cal U}^{(ge)}(\tau)|\Psi_{B}\rangle\langle\Psi_{B}|\mu^{2}|g\rangle\langle g|.\label{eq:Wg_2}\end{equation}
The superoperator ${\cal U}^{(eg)}(t)$ is the evolution superoperator
of a coherence projector $\sum_{n}|e_{n}\rangle\langle g|$ and analogically
for ${\cal U}^{(ge)}(t)$. After a delay $T$ the third ultrafast
pulse is applied and the non-linear signal is recorded. The signal
corresponds to indirectly to non-linear polarization of the sample,
and is usually measured in frequency domain\[
E^{(3)}(\omega,T,\tau)\approx i\ tr\{\mu{\cal U}^{(eg)}(\omega)\mu\]
\begin{equation}
\times({\cal U}^{(e)}(T){\cal U}^{(ge)}(\tau)\mu\rho_{g}\mu+{\cal U}^{(g)}(T){\cal U}^{(ge)}(\tau)\mu^{2}\rho_{g})W_{eq}\}.\label{eq:E3}\end{equation}
Here, we denoted\begin{equation}
{\cal U}^{(eg)}(\omega)=\int\limits _{0}^{\infty}dt\ e^{i\omega t}{\cal U}^{(eg)}(t),\label{eq:U_omega}\end{equation}
and $\rho_{g}=|g\rangle\langle g|$. In 2D coherent spectroscopy,
the signal is in addition Fourier transformed along the time delay
$\tau$, so that the spectrum is defined as\begin{equation}
S_{2D}(\omega_{t},T,\omega_{\tau})=\int\limits _{-\infty}^{\infty}d\tau\ e^{-i\omega_{\tau}\tau}E^{(3)}(\omega_{t},T,\tau).\label{eq:2D_spect}\end{equation}
The spectrum defined in this way has a suitable interpretation of
an absorption -- absorption and absorption -- stimulated emission
correlation plot, with a different waiting times $T$ between the
two events. The 2D spectrum is in practice measured with finite pulses,
and the measured time domain signal is thus a triple convolution of
the responses to a delta pulse excitation, with the actual finite
pulses \cite{Brixner2005b}.
From this rough sketch of the principles and the information content
of the coherent 2D spectroscopy it should be clear that 2D spectroscopy
is aimed at disentangling the dynamics of the system during the time
delay $T$. In the so-called Markov approximation, when the dynamics
in time intervals $\tau$, $T$ and $t$ is assumed separable, and
the bath is assumed stationary, the ground state evolution during
interval $T$ can be neglected. Then 2D measurement essentially accesses
the reduced evolution superoperator, Eq. (\ref{eq:Ue_red}) and possibly
also the more general superoperator \begin{equation}
\bar{{\cal U}}^{(e)}(t,\tau)=tr_{B}\{{\cal U}^{(e)}(t)U^{(ge)}(\tau)W_{eq}\}.\label{eq:Ue_ttau}\end{equation}
We will show below that this superoperator, together with the light
properties, determines the way in which the molecule is excited in
a general case, even at illumination by natural light.
\section{Excitation by Light\label{sec:Excitation-by-Thermal}}
In order to account for general light properties we will consider
the problem fully quantum mechanically, and assume only deterministic
evolution of the system wavefunction. The Hamiltonian of the system
reads\[
H=H_{M}+H_{B}+H_{R}+H_{S}\]
\begin{equation}
+H_{M-B}+H_{M-R}+H_{M-S}+H_{B-S}+H_{B-R}+H_{R-S}.\label{eq:Ham_compl}\end{equation}
We have divided the system into a molecule ($H_{M}$), its environment
or bath ($H_{B}$), the radiation ($H_{R}$) and the light emitting
body (LEB) which produces it, e.g. Sun or laser medium ($H_{S}$).
It seems reasonable to neglect a direct interaction between the molecule
(together with its environment) and the molecules of the LEB. Consequently,
the terms $H_{M-S}$ and $H_{B-S}$ can be disregarded. To make the
treatment simpler we can also neglect the interaction between radiation
and the molecular environment, $H_{B-R}$. The assumption is that
the energy of the molecular transition that is used to harvest light
for photosynthetic purposes is much larger than any of the transitions
in this environment and the two regions of the light spectrum can
thus be treated separately. One can also assume that the part of radiation
spectrum which would interact with the bath is simply filtered out,
and the environment is kept at certain temperature by other means.
\subsection{Radiation entangled with the light emitting body}
An important special case is the one in which the radiation and the
LEB is in equilibrium with each other so that the radiation is described
by the canonical equilibrium density matrix\begin{equation}
W_{R}^{(eq)}=\sum_{\lambda\bm{q}}\frac{e^{-\frac{N_{\lambda\bm{q}}\hbar\omega_{\bm{q}}}{k_{B}T}}}{Z_{\lambda\bm{q}}}|N_{\lambda\bm{q}}\rangle\langle N_{\lambda\bm{q}}|.\label{eq:R_eq}\end{equation}
Here, $|N_{\lambda\bm{q}}\rangle$ is the $N$-photon state of the
radiation mode with polarization vector $\bm{e}_{\lambda}$ and wave-vector
$\bm{q}$. As we have already noted above, the statistical concept
of density matrix will be replaced here with the concept of entangled
states, so that we can describe the whole system by its state vector.
Thus, we introduce a state vector \begin{equation}
|\Xi(t)\rangle=\sum_{\lambda\bm{q}}\sum_{N_{\lambda\bm{q}}}c_{N\lambda\bm{q}}(t)|N_{\lambda\bm{q}}\rangle|\phi_{N\lambda\bm{q}}(t)\rangle,\label{eq:Xi}\end{equation}
in which the light is fully entangled with the states $|\phi_{N\lambda\bm{q}}(t)\rangle$
of the LEB . The LEB states have to fulfill the condition \begin{equation}
\langle\phi_{N\lambda\bm{q}}(t)|\phi_{N^{\prime}\lambda^{\prime}\bm{q}^{\prime}}(t)\rangle=\delta_{\lambda\lambda^{\prime}}\delta_{\bm{q}\bm{q}^{\prime}}\delta_{N_{\lambda\bm{q}}N_{\lambda\bm{q}}^{\prime}},\label{eq:condition_phi}\end{equation}
so that when the total density matrix of the LEB and radiation is
averaged over the states of the body, we obtain Eq. (\ref{eq:R_eq}).
$W_{R}^{(eq)}$ is recovered provided that \begin{equation}
|c_{N\lambda\bm{q}}(t)|^{2}=\frac{e^{-\frac{N_{\lambda\bm{q}}\hbar\omega_{\bm{q}}}{k_{B}T}}}{Z_{\lambda\bm{q}}}.\label{eq:coeffs_c}\end{equation}
In the absence of the light absorbing body, the evolution of the state
$|\Xi(t)\rangle$ is governed by the Hamiltonian\begin{equation}
H_{L}=H_{R}+H_{S}+H_{R-S},\label{eq:HL}\end{equation}
and \[
U_{L}(t)=\Theta(t-t_{0})\]
\begin{equation}
\times\exp\left\{ -\frac{i}{\hbar}(H_{S}+H_{R}+H_{R-S})(t-t_{0})\right\} \label{eq:UL}\end{equation}
is the corresponding evolution operator.
\subsection{Equation of motion}
For the subsequent treatment of the system dynamics, we introduce
the interaction picture with respect to Hamiltonian operators $H_{M}$,
$H_{B}$ and $H_{L}$,\begin{equation}
H^{(I)}(t)=H_{M-B}(t)+H_{M-R}(t),\label{eq:H_I}\end{equation}
where\begin{equation}
H_{M-B}(t)=U_{M}^{\dagger}(t)U_{B}^{\dagger}(t)H_{M-B}U_{B}(t)U_{M}(t),\label{eq:HMBt}\end{equation}
and \begin{equation}
H_{M-R}(t)=U_{M}^{\dagger}(t)U_{L}^{\dagger}(t)H_{M-R}U_{L}(t)U_{M}(t).\label{eq:HMRt}\end{equation}
Equation of motion for the total state vector in the interaction picture
\begin{equation}
|\Psi^{(I)}(t)\rangle=U_{M}^{\dagger}(t)U_{B}^{\dagger}(t)U_{L}^{\dagger}(t)|\Psi(t)\rangle\label{eq:I_picture}\end{equation}
thus reads\[
\frac{\partial}{\partial t}|\Psi^{(I)}(t)\rangle=\]
\begin{equation}
-\frac{i}{\hbar}\left(H_{M-B}(t)+H_{M-R}(t)\right)|\Psi^{(I)}(t)\rangle.\label{eq:Sch_I}\end{equation}
The solution of Eq. (\ref{eq:Sch_I}) can be found formally as\[
|\Psi^{(I)}(t)\rangle=\exp_{+}\Big\{-\frac{i}{\hbar}\int\limits _{t_{0}}^{t}d\tau(H_{M-B}(\tau)\]
\begin{equation}
+H_{M-R}(\tau))\Big\}|\Psi_{0}\rangle.\label{eq:solution_Psi_I}\end{equation}
We will assume that the system is initially in the state $|\Psi_{0}\rangle$
of Eq. (\ref{eq:state_ini}). With this choice we have\begin{equation}
H_{M-B}(t)|\Psi_{0}\rangle=0.\label{eq:HMB_on_Psi_0}\end{equation}
Further in this paper, we will assume weak interaction with the radiation,
so that it can be described by linear theory. Thus, we need to collect
all terms in the expansion of Eq. (\ref{eq:solution_Psi_I}) which
include one occurrence of $H_{M-R}(t)$. Thanks to Eq. (\ref{eq:HMB_on_Psi_0}),
however, all terms where $H_{M-R}(t)$ is not on the far right of
the expression are equal to zero. Eq. (\ref{eq:solution_Psi_I}) therefore
simplifies into a series\[
|\Psi^{(I)}(t)\rangle=|\Psi_{0}\rangle-\frac{i}{\hbar}\int\limits _{t_{0}}^{t}d\tau H_{M-R}(\tau)|\Psi_{0}\rangle\]
\begin{equation}
-\frac{1}{\hbar^{2}}\int\limits _{t_{0}}^{t}d\tau\int\limits _{t_{0}}^{t}d\tau^{\prime}H_{M-B}(\tau)H_{M-R}(\tau^{\prime})|\Psi_{0}\rangle+\dots\ .\label{eq:sol_series}\end{equation}
Now we introduce a projector ${\cal P}_{e}$ that excludes the excitonic
ground state $|g\rangle$\begin{equation}
{\cal P}_{e}=\sum_{n}|e_{n}\rangle\langle e_{n}|.\label{eq:P_e}\end{equation}
Applying this projector to Eq. (\ref{eq:sol_series}) has only the
effect of eliminating the first term of the series. Introducing abbreviations
\begin{equation}
|S(t)\rangle=-\frac{i}{\hbar}\int\limits _{t_{0}}^{t}d\tau H_{M-R}(\tau)|\Psi_{0}\rangle\label{eq:Source_state}\end{equation}
and\begin{equation}
|\Psi_{e}^{(I)}(t)\rangle={\cal P}_{e}|\Psi^{(I)}(t)\rangle,\label{eq:Psi_I_e}\end{equation}
we can write\[
|\Psi_{e}^{(I)}(t)\rangle=|S(t)\rangle-\frac{i}{\hbar}\int\limits _{t_{0}}^{t}d\tau H_{M-B}(\tau)|S(\tau)\rangle\]
\begin{equation}
-\frac{1}{\hbar^{2}}\int\limits _{t_{0}}^{t}d\tau\int\limits _{t_{0}}^{\tau}d\tau^{\prime}H_{M-B}(\tau)H_{M-B}(\tau^{\prime})|S(\tau^{\prime})\rangle+\dots\ .\label{eq:P_e_series}\end{equation}
It is possible to verify easily that this series is a solution of
the equation \begin{equation}
\frac{\partial}{\partial t}|\Psi_{e}^{(I)}(t)\rangle=-\frac{i}{\hbar}H_{M-B}(t)|\Psi_{e}^{(I)}(t)\rangle+\frac{\partial}{\partial t}|S(t)\rangle,\label{eq:source_eq}\end{equation}
with initial condition $|\Psi_{e}^{(I)}(t_{0})\rangle=0$.
\subsection{Pumping source term}
Eq. (\ref{eq:source_eq}) is an equation of motion for the excited
states of an excitonic aggregate pumped by a source term \begin{equation}
|S^{\prime}(t)\rangle=\frac{\partial}{\partial t}|S(t)\rangle=-\frac{i}{\hbar}H_{M-R}(t)|\Psi_{0}\rangle.\label{eq:S_prime}\end{equation}
Hamiltonian $H_{M-R}$ will be assumed in the dipole approximation,
i.e.\begin{equation}
H_{M-R}\approx-\bm{\mu}\cdot\bm{E}_{T}(\bm{r}),\label{eq:H_MR_form}\end{equation}
where $\bm{\mu}$ is the transition dipole moment operator of the
aggregate\begin{equation}
\bm{\mu}=\sum_{n}\bm{d}_{n}|e_{n}\rangle\langle g|+h.c.,\label{eq:mu_def}\end{equation}
and $\bm{E}_{T}$ is the operator of the (transversal) electric field
of the radiation\[
\bm{E}_{T}(\bm{r})=-i\sum_{\lambda\bm{q}}\Big(\bm{e}_{\lambda-\bm{q}}f_{-\bm{q}}(\bm{r})a_{\lambda\bm{q}}^{\dagger}\]
\begin{equation}
-\bm{e}_{\lambda\bm{q}}f_{\bm{q}}(\bm{r})a_{\lambda\bm{q}}\Big),\label{eq:E_T_def}\end{equation}
with \begin{equation}
f_{\bm{q}}(\bm{r})=\sqrt{\frac{\hbar\omega_{\bm{q}}}{2\epsilon_{0}\Omega}}e^{i\bm{q}\cdot\bm{r}}.\label{eq:flq}\end{equation}
Here, $\Omega$ is a quantization volume.
We consider a molecule much smaller than the wavelength of the light,
so that $e^{i\bm{q}\cdot\bm{r}}$ is constant in the volume of the
molecule. The origin of the coordinates can thus be conveniently put
into the molecule yielding $e^{i\bm{q}\cdot\bm{r}}\approx1$. The
interaction Hamiltonian in Eq. (\ref{eq:S_prime}) then reads\[
H_{M-R}(t)=i\sum_{\lambda\bm{q}}\mu_{\lambda\bm{q}}(t)f_{\lambda-\bm{q}}(0)a_{\lambda\bm{q}}^{\dagger}(t)\]
\begin{equation}
-i\mu_{\lambda\bm{q}}(t)f_{\lambda\bm{q}}(0)a_{\lambda\bm{q}}(t),\label{eq:HMR_in_a}\end{equation}
where the creation and annihilation operators of the field are in
the interaction picture with respect to Hamiltonian $H_{L}$, i.e.\begin{equation}
a_{\lambda\bm{q}}^{\dagger}(t)=U_{L}^{\dagger}(t)a_{\lambda\bm{q}}^{\dagger}U_{L}(t),\label{eq:adagger_t}\end{equation}
\begin{equation}
a_{\lambda\bm{q}}(t)=U_{L}^{\dagger}(t)a_{\lambda\bm{q}}U_{L}(t).\label{eq:a_t}\end{equation}
The transition dipole moment operator projected on the polarization
vector of a mode $\lambda\bm{q}$ appears in the interaction picture
with respect to Hamiltonian $H_{M}$,\begin{equation}
\mu_{\lambda\bm{q}}(t)=U_{M}^{\dagger}(t)\bm{\mu}\cdot\bm{e}_{\lambda\bm{q}}U_{M}(t).\label{eq:mu_lq}\end{equation}
The evolution operator $U_{L}(t)$, Eq. (\ref{eq:UL}), can be rewritten
as\begin{equation}
U_{L}(t)=U_{S}(t)U_{R}(t){\cal U}_{R-S}(t),\label{eq:UL_SR}\end{equation}
where\[
U_{R-S}(t)=\Theta(t-t_{0})\exp_{+}\Big\{-\frac{i}{\hbar}\int\limits _{t_{0}}^{t}d\tau U_{S}^{\dagger}(\tau)U_{R}^{\dagger}(\tau)\]
\begin{equation}
\times H_{R-S}U_{R}(\tau)U_{S}(\tau)\Big\}.\label{eq:calU_RS}\end{equation}
Since Hamiltonian $H_{S}$ commutes with the radiation operators and
\begin{equation}
U_{R}^{\dagger}(t)a_{\lambda\bm{q}}U_{R}(t)=e^{-i\omega_{\bm{q}}t}a_{\lambda\bm{q}},\label{eq:UaU}\end{equation}
we have\begin{equation}
a_{\lambda\bm{q}}^{\dagger}(t)=\tilde{a}_{\lambda\bm{q}}^{\dagger}(t)e^{i\omega_{\bm{q}}t},\label{eq:adag}\end{equation}
and \begin{equation}
a_{\lambda\bm{q}}(t)=\tilde{a}_{\lambda\bm{q}}(t)e^{-i\omega_{\bm{q}}t}.\label{eq:a}\end{equation}
Here, we introduced slow oscillating envelops \begin{equation}
\tilde{a}_{\lambda\bm{q}}^{\dagger}(t)=U_{R-S}^{\dagger}(t)a_{\lambda\bm{q}}^{\dagger}U_{R-S}(t),\label{eq:adag_til}\end{equation}
and\begin{equation}
\tilde{a}_{\lambda\bm{q}}(t)=U_{R-S}^{\dagger}(t)a_{\lambda\bm{q}}U_{R-S}(t).\label{eq:a_til}\end{equation}
Inserting these expressions into Eq. (\ref{eq:HMR_in_a}) we can distinguish
two terms associated with the transition from the ground state $|g\rangle$
to an excited state $|e_{a}\rangle$ with respective phase factors
$e^{i(\omega_{ag}-\omega_{\bm{q}})t}$ and $e^{i(\omega_{ag}+\omega_{\bm{q}})t}$.
While the first one will lead to a resonance excitation around $\omega_{\bm{q}}\approx\omega_{ag}$,
the later term is oscillating fast and will therefore contribute very
little compared to the former one. Thus we drop the fast oscillating
part, and obtain the source term in the form\[
|S^{\prime}(t)\rangle=\frac{1}{\hbar}\sum_{\lambda\bm{q}}\mu_{\lambda\bm{q}}(t)f_{\lambda\bm{q}}\]
\begin{equation}
\times a_{\lambda\bm{q}}(t)|g\rangle|\Phi_{B}\rangle|\Xi_{0}\rangle.\label{eq:S_prime_derived}\end{equation}
Using this form of the source term, we can find state into which the
molecule is weakly driven by any type of light.
\subsection{Excited state dynamics under pumping}
So far we have treated the problem systematically using the wavefunction
approach. The time evolution of the system wavefunction is governed
by Eq. (\ref{eq:Sch_I}). To find the probabilities of creating population
on and coherence between certain excitonic levels $|e_{a}\rangle$
we solve Eq. (\ref{eq:Sch_I}) formally,\begin{equation}
|\Psi_{e}^{(I)}(t)\rangle=\int\limits _{t_{0}}^{t}d\tau U_{M-B}(t-\tau)|S^{\prime}(\tau)\rangle.\label{eq:solution_psi_e}\end{equation}
Here, we used the fact that $|\Psi_{e}(t_{0})\rangle=0$. Now let
us evaluate matrix element $P_{ab}(t)=\langle\Psi_{e}(t)|{\cal P}_{ab}|\Psi_{e}(t)\rangle$
of a projector \begin{equation}
{\cal P}_{ab}=|e_{a}\rangle\langle e_{b}|,\label{eq:proj_nm}\end{equation}
which gives the probability of finding the molecule in state $|e_{a}\rangle$
if $a=b$, or characterizes the amount of coherence between states
$|e_{a}\rangle$ and $|e_{b}\rangle$ if $a\neq b$. Note that we
have removed the interaction picture, Eq. (\ref{eq:I_picture}). We
have\[
P_{ab}(t)=\frac{1}{\hbar^{2}}\int\limits _{t_{0}}^{t}d\tau\int\limits _{t_{0}}^{t}d\tau^{\prime}\sum_{\lambda\bm{q},\lambda^{\prime}\bm{q}^{\prime}}f_{\lambda\bm{q}}(f_{\lambda^{\prime}\bm{q}^{\prime}})^{*}\]
\[
\times\langle\Xi_{0}|\tilde{a}_{\lambda\bm{q}}^{\dagger}(\tau)\tilde{a}_{\lambda^{\prime}\bm{q}^{\prime}}(\tau^{\prime})|\Xi_{0}\rangle\]
\begin{equation}
\times\langle e_{b}|\bar{{\cal U}}^{(e)}(t-\tau,\tau-\tau^{\prime})\rho_{\lambda\bm{q},\lambda^{\prime}\bm{q}^{\prime}}^{0}|e_{a}\rangle,\label{eq:Pab_general}\end{equation}
where the evolution superoperator $\bar{{\cal U}}^{(e)}(t-\tau,\tau-\tau^{\prime})$
has been defined in Eq. (\ref{eq:Ue_ttau}).
In Eq. (\ref{eq:Pab_general}), the light is represented by a first
order correlation function\[
I_{\lambda\bm{q},\lambda^{\prime}\bm{q}}^{(1)}(\tau,\tau^{\prime})=(f_{\lambda\bm{q}}(f_{\lambda^{\prime}\bm{q}^{\prime}})^{*})^{-1}\]
\begin{equation}
\times\langle\Xi_{0}|\tilde{a}_{\lambda\bm{q}}^{\dagger}(\tau)\tilde{a}_{\lambda^{\prime}\bm{q}^{\prime}}(\tau^{\prime})|\Xi_{0}\rangle\label{eq:I1_def}\end{equation}
(see e. g. Ref. \cite{LoundonBook}), which comprises all its relevant
properties. We also denoted \begin{equation}
\rho_{\lambda\bm{q},\lambda^{\prime}\bm{q}^{\prime}}^{0}=\frac{1}{\hbar^{2}}\mu_{\lambda\bm{q}}|g\rangle\langle g|\mu_{\lambda^{\prime}\bm{q}^{\prime}}.\label{eq:rho_ini}\end{equation}
The quantities $P_{ab}(t)$ are the matrix elements of the RDM ($P_{ab}(t)=\langle e_{b}|\rho(t)|e_{a}\rangle$)
of the system which reads\[
\rho(t)=\int\limits _{t_{0}}^{t}d\tau\int\limits _{t_{0}}^{t}d\tau^{\prime}\bar{{\cal U}}^{(e)}(t-\tau,\tau-\tau^{\prime})\]
\begin{equation}
\times\sum_{\lambda\bm{q},\lambda^{\prime}\bm{q}^{\prime}}\rho_{\lambda\bm{q},\lambda^{\prime}\bm{q}^{\prime}}^{0}I_{\lambda\bm{q},\lambda^{\prime}\bm{q}}^{(1)}(\tau,\tau^{\prime}).\label{eq:rho_final}\end{equation}
For a weakly driven system, Eq. (\ref{eq:rho_final}) has a very wide
range of applicability. We will discuss its application to thermal
light and pulsed coherent light in the following section.
\section{Discussion\label{sec:Discussion}}
Thorough discussion of excitation dynamics of molecular systems excited
by incoherent light was made in Ref. \cite{Brumer}. Molecular systems
were considered without the bath effect which is however significant
for light harvesting. Eq. (\ref{eq:rho_final}) contains reduced evolution
superoperator of the molecular system so that the state of the system
created by the incident light depends on its reduced dynamics. It
is not possible to consider a general case of such dynamics analytically,
and we will therefore commit ourselves to some simple cases.
In so called secular and Markov approximations (see e.g. Ref. \cite{MayKuehnBook})
matrix elements of the evolution superoperator governing the coherences
take a very simple form. First, it is possible to separate the two
time arguments in the superoperator $\bar{{\cal U}}^{(e)}(t,\tau)$
so that \begin{equation}
\bar{{\cal U}}^{(e)}(t,\tau)=\bar{{\cal U}}^{(e)}(t)\bar{{\cal U}}^{(eg)}(\tau).\label{eq:Markov_U}\end{equation}
Since each coherence is independent of the population dynamics and
of other coherences, the one-argument superoperator elements read\begin{equation}
\bar{{\cal U}}_{abab}^{(e)}(t)=e^{-i\omega_{ab}t-(\Gamma_{a}+\Gamma_{b})t},\label{eq:Uabab}\end{equation}
and\begin{equation}
\bar{{\cal U}}_{agag}^{(eg)}(t)=e^{-i\omega_{ag}t-\Gamma_{a}t}.\label{eq:Uagag}\end{equation}
Here the dephasing rate \begin{equation}
\Gamma_{a}=\gamma_{p}+\frac{1}{2}K_{a}\label{eq:Gamma_ab}\end{equation}
comprises the pure dephasing rate $\gamma_{p}$ and the rate $K_{a}$
of depopulation, i.e. the sum of transition rates from state $|e_{a}\rangle$
to other states. A simplified treatment of the populations is possible
for the states that are only depopulated, i.e. no contributions to
the population can be attributed to the transfer from other levels.
They are found at the top of the energetic funnel of the antenna.
For these states we have\begin{equation}
\bar{{\cal U}}_{aaaa}^{(e)}(t)=e^{-K_{a}t}.\label{eq:Uaaaa}\end{equation}
Eqs. (\ref{eq:Uabab}) to (\ref{eq:Uaaaa}) neglect all coherence
transfer effects, as well as possible coupling between the dynamics
of population and coherence.
\subsection{Excitation of coherences by thermal light}
For an equilibrium thermal light the correlation function $I_{\lambda\bm{q},\lambda^{\prime}\bm{q}^{\prime}}^{(1)}(\tau,\tau^{\prime})$
depends only on the difference of the times $\tau$ and $\tau^{\prime}$.
As discussed above, $|\Xi_{0}\rangle$ represents the equilibrium
of the system described by Hamiltonian $H_{L}$. The equilibrium density
matrix is stationary, i.e.\begin{equation}
U_{L}(t)|\Xi_{0}\rangle\langle\Xi_{0}|U_{L}^{\dagger}(t)=|\Xi_{0}\rangle\langle\Xi_{0}|,\label{eq:dens_eq}\end{equation}
so we can write\[
I_{\lambda\bm{q},\lambda^{\prime}\bm{q}}^{(1)}(\tau,\tau^{\prime})=|f_{\lambda\bm{q}}|^{-2}\langle\Xi_{0}|\tilde{a}_{\lambda\bm{q}}^{\dagger}(\tau-\tau^{\prime})\tilde{a}_{\lambda\bm{q}}(0)|\Xi_{0}\rangle\]
\[
\times e^{i\omega_{\bm{q}}(\tau-\tau^{\prime})}\delta_{\lambda\lambda^{\prime}}\delta_{\bm{q}\bm{q}^{\prime}}\]
\begin{equation}
\equiv|f_{\lambda\bm{q}}|^{-2}\tilde{I}_{\lambda\bm{q}}(\tau-\tau^{\prime})\delta_{\lambda\lambda^{\prime}}\delta_{\bm{q}\bm{q}^{\prime}}.\label{eq:n_def}\end{equation}
It can be shown that \begin{equation}
\tilde{I}_{\lambda\bm{q}}(-t)=\tilde{I}_{\lambda\bm{q}}^{*}(t).\label{eq:n_sym}\end{equation}
Assuming some simple form of a light correlation function, e.g. \begin{equation}
\tilde{I}_{\lambda\bm{q}}(t)=I_{\lambda\bm{q}}^{0}e^{-\frac{|t|}{\tau_{d}}+i\omega_{\bm{q}}t},\label{eq:I_collisional}\end{equation}
we obtain for the populations\[
\rho_{aa}(t)=2Re\sum_{\lambda\bm{q}}I_{\lambda\bm{q}}^{0}[\rho_{\lambda\bm{q}}]_{aa}\int\limits _{t_{0}}^{t}d\tau\int\limits _{t_{0}}^{\tau}d\tau^{\prime}e^{-K_{a}(t-\tau)}\]
\begin{equation}
\times e^{-\Gamma_{a}(\tau-\tau^{\prime})-\frac{\tau-\tau^{\prime}}{\tau_{d}}-i(\omega_{ag}-\omega_{\bm{q}})(\tau-\tau^{\prime})}\label{eq:pop_integral_form}\end{equation}
Here, $[\rho]_{ab}\equiv\langle e_{a}|\rho|e_{b}\rangle$. We utilized
Eq. (\ref{eq:I_collisional}) and the fact that, by definition (see
Eqs. (\ref{eq:I1_def}) and (\ref{eq:rho_ini})), the time $\tau$
corresponds to the action of the dipole moment operator from left,
whereas time $\tau^{\prime}$ corresponds to the same action from
the right. At long times $t-t_{0}\rightarrow\infty$ this yields \begin{equation}
\rho_{aa}^{\infty}=\sum_{\lambda\bm{q}}\frac{2}{K_{a}}\frac{(\Gamma_{a}+\frac{1}{\tau_{d}})I_{\lambda\bm{q}}^{0}[\rho_{\lambda\bm{q}}^{0}]_{aa}}{(\omega_{ag}-\omega_{\bm{q}})^{2}+(\Gamma_{a}+\frac{1}{\tau_{d}})^{2}}.\label{eq:rho_aa_inf}\end{equation}
However, neglecting the influence of environment as in Ref. \cite{Brumer}
yields\begin{equation}
\rho_{aa}^{long}(t-t_{0})=\sum_{\lambda\bm{q}}\frac{\tau_{d}^{-1}2I_{\lambda\bm{q}}^{0}[\rho_{\lambda\bm{q}}^{0}]_{aa}(t-t_{0})}{(\omega_{ag}-\omega_{\bm{q}})^{2}+\frac{1}{\tau_{d}^{2}}},\label{eq:rho_aa_delta}\end{equation}
which grows linearly with time.
For coherences we have\[
\rho_{ab}^{\infty}=2\sum_{\lambda\bm{q}}I_{\lambda\bm{q}}^{0}[\rho_{\lambda\bm{q}}^{0}]_{ab}\frac{1}{i\omega_{ab}+(\Gamma_{a}+\Gamma_{b})}\]
\[
\times\Big[\frac{1}{i(\omega_{ag}-\omega_{\bm{q}})+\Gamma_{a}+\frac{1}{\tau_{b}}}\]
\begin{equation}
+\frac{1}{-i(\omega_{bg}-\omega_{\bm{q}})+\Gamma_{b}+\frac{1}{\tau_{b}}}\Big],\label{eq:rho_ab_inf}\end{equation}
which turns into Eq. (\ref{eq:rho_aa_inf}) for $a=b$ (with additional
assumption $K_{a}=2\Gamma_{a}$). In case of no dephasing, the first
fraction in Eq. (\ref{eq:rho_ab_inf}) yields a delta function $\delta(\omega_{ab})$
\cite{Brumer}. Thus, for slow or non-existent relaxation due to interaction
with environment, the system is excited predominantly into a state
represented by a diagonal RDM, as all coherence terms are negligible
compared to the linearly growing population. For fast relaxation,
the coherences may be of the same order of magnitude as the populations.
The case of very fast relaxation is particularly interesting. It was
suggested previously that coherent dynamics can be relevant for the
\emph{in vivo} case, because the fluctuating light from the Sun corresponds
to a train of ultrafast spikes \cite{Cheng-Review}. The relaxation
of the antenna must be in such a case fast enough to prevent averaging
over many such spikes. Eqs. (\ref{eq:rho_aa_inf}) and (\ref{eq:rho_ab_inf})
with large $K_{a}$ describe just such a situation. The RDM created
by incoherent light resembles in certain sense the one created by
ultrafast pulses; it represents a linear combination of excitons.
The coherences in Eq. (\ref{eq:rho_ab_inf}) are however static at
long times.
In our demonstration we concentrated on a simple model assuming both
Markov and secular approximations to be valid. The presence or absence
of coherences has no significance in such a case, and more involved
theories of the RDM dynamics \cite{Ishizaki1,Ishizaki2,Olsina2010a}
have to be used to investigate the role of coherences in energy transfer
processes by Eq. (\ref{eq:rho_final}).
\subsection{Coherent pulsed light}
In derivation of Eq. (\ref{eq:rho_final}) we assumed certain initial
state $|\Xi_{0}\rangle$ of the system composed of the light and its
source. The condition that the light is in a stationary state, fully
entangled with its source, has only been used to simplify the correlation
function $I_{\lambda\bm{q},\lambda^{\prime}\bm{q}^{\prime}}^{(1)}(\tau,\tau^{\prime})$
for the case of the thermal light. In a general case $|\Xi_{0}\rangle$
will not represent an equilibrium state. It can indeed describe even
systems such as a laser producing coherent Gaussian light pulses with
some carrier frequency $\omega_{0}$ and a width parameter $\Delta$.
If we in addition assume that the light is described by a single
polarization, and that we consider the dynamics after one such pulse
centered at time $t=\tau_{0}$, the light is described as\begin{equation}
\sum_{\lambda\bm{q},\lambda^{\prime}\bm{q}^{\prime}}I_{\lambda\bm{q}\lambda^{\prime}\bm{q}^{\prime}}^{(1)}(\tau,\tau^{\prime})=I_{0}e^{-\frac{(\tau-\tau_{0})^{2}}{\Delta^{2}}-\frac{(\tau^{\prime}-\tau_{0})^{2}}{\Delta^{2}}}.\label{eq:Gauss_exc}\end{equation}
Coherence element created by such light reads\[
\rho_{ba}(t)=e^{i\omega_{ab}t}\int\limits _{t_{0}}^{t}d\tau\int\limits _{t_{0}}^{t}d\tau^{\prime}e^{-i(\omega_{ag}-\omega_{\bm{q}})\tau}e^{i(\omega_{bg}-\omega_{\bm{q}})\tau^{\prime}}\]
\begin{equation}
\times e^{-\Gamma_{a}(t-\tau)-\Gamma_{b}(t-\tau^{\prime})}I_{0}e^{-\frac{(\tau-\tau_{0})^{2}}{\Delta^{2}}-\frac{(\tau^{\prime}-\tau_{0})^{2}}{\Delta^{2}}}\rho_{ba}^{0},\label{eq:Pab_model_coh}\end{equation}
where $\rho_{ba}^{0}=\frac{1}{\hbar^{2}}\langle e_{b}|\mu|g\rangle\langle g|\mu|e_{a}\rangle$.
In the limit of ultrashort pulses when $e^{-\frac{(\tau-\tau_{0})^{2}}{\Delta^{2}}}\rightarrow\alpha\delta(\tau-\tau_{0})$
the pulse creates a pure state at $\tau_{0}$, which then dephases
as\[
\rho_{ba}(t)=\Theta(t-\tau_{0})e^{-(\Gamma_{a}+\Gamma_{b})(t-\tau_{0})}\]
\begin{equation}
\times e^{i\omega_{ab}(t-\tau_{0})}\rho_{ba}^{0}I_{0}\alpha^{2}.\label{eq:rho_ba_time}\end{equation}
In case of a finite pulse and no dephasing our results coincides with
those found in Ref. \cite{Brumer}.
\subsection{Mediated excitation}
The major difference between excitation by the thermal light and a
coherence pulse is in the occurrence of a sudden event which populates
a nearly pure state of the excited state band. Clearly, a single molecule
interacting with an ideal continuum of radiation modes in equilibrium
does not experience such sudden events. Rather, its interaction with
light corresponds to a continuous pumping, and the suddenness of the
photon arrival is the consequence of our ability to register only
classical outcomes. In order to register them we have to interact
with the system and become entangled with it. Our experience is that
macroscopic systems interacting with low intensity light can be used
to detect single photons, and certain more or less definite times
can be attributed to their arrivals. Interaction of a photon with
a macroscopic detector yields a temporal localization of the arrival
event. A mesoscopic system may play a role of such a detector (mediator)
that provides its fluctuations to be harvested by dedicated nano sized
antenna. Green photosynthetic bacteria, from which the photosynthetic
complex FMO was isolated, collect light mainly by means of so-called
chlorosoms \cite{Blankenship,Chlorosom}. The chlorosom is a self-assembled
aggregate of $\sim10^{5}$ BChls and carotenoids with very little
protein. The typical dimensions the chlorosom are of the order of
$100$ nm \cite{Chlorosom}. It does not seem to be organized as an
energy funnel \cite{Nozawa94,Psencik04}, and the energy transfer
time between its main body and the base plate to which FMO complexes
are attached is of the order of $120$ ps \cite{Psencik03}, i.e rather
slow. The excitation in such a mesoscopic system may have enough time
to become localized through interaction with the large number of the
systems DOF and arrive at the FMO complex in a particle like, i.e.
also temporally localized fashion.
In this section, we will generalize our result, Eq. (\ref{eq:rho_final}),
for a case when the excitation of the photosynthetic systems occurs
by transfer from another system. We will therefore assume that our
molecule does not interact directly with light, but is pumped in a
similar fashion by another system. The source term, Eq. (\ref{eq:S_prime_derived}),
is then generalized as\[
|S^{\prime}(t)\rangle=\frac{i}{\hbar}A(t)|g\rangle|\Phi_{B}\rangle\]
\begin{equation}
\times\left(\sum_{n}\alpha_{n}(t)|\xi_{n}\rangle|\phi_{n}(t)\rangle\right).\label{eq:S_prime_gene}\end{equation}
Here, $A=\sum_{\alpha,n}|e_{n}\rangle|\xi_{g}\rangle\langle\xi_{\alpha}|\langle g|+h.c.$
is the molecule--mediator interaction Hamiltonian and the time dependence
results from the interaction picture\begin{equation}
A(t)=U_{M}^{\dagger}(t)U_{A}^{\dagger}(t)AU_{A}(t)U_{M}(t).\label{eq:A_op}\end{equation}
We denoted the ground and excited states of the mediator by $|\xi_{g}\rangle$
and $|\xi_{n}\rangle$, respectively. The state of the molecule at
long times is in analogy with Eq. (\ref{eq:rho_final})\[
\rho(t)=\int\limits _{t_{0}}^{t}d\tau\int\limits _{t_{0}}^{t}d\tau^{\prime}{\cal U}^{(e)}(t-\tau,\tau-\tau^{\prime})\sum_{nn^{\prime}}\alpha_{n}^{*}(\tau)\alpha_{n^{\prime}}(\tau)\]
\begin{equation}
\times\langle\xi_{n}|A(\tau)A(\tau^{\prime})|\xi_{n^{\prime}}\rangle\langle\phi_{n}(\tau)|\phi_{n^{\prime}}(\tau^{\prime})\rangle.\label{eq:rho_fina_mosol}\end{equation}
The complicated two-point correlation function in Eq. (\ref{eq:rho_fina_mosol})
results from the pumping of the mediator similarly to the direct pumping
of the molecule in Eq. (\ref{eq:rho_final}). A mesoscopic system
especially when excited will, however, always exhibit fluctuations
which will prevent the correlation function from having a simple smooth
dependence without recurrences. Such recurrences can temporally localize
the excitation events of the molecule. In such an excitation regime,
when coherent dynamics from different excitation times do not interfere,
optimization of the FMO's energy channeling capability for case of
initially coherent states would be an advantage.
\subsection{Outlook }
More research into specific forms of both the light correlation function
for different situation that may occur \emph{in vivo}, and the analogical
interaction of systems like FMO with mesoscopic antennae is clearly
needed.\emph{ }Ultrafast spectroscopic experiments play a pivotal
role in this research by yielding information about the system response
to the light. To conclude on the utility of coherent dynamics for
the function of the photosynthetic system is, however, only possible
by taking into account the properties of light at the natural conditions,
for which the results of this paper provide means. If the coherent
dynamics observed in some photosynthetic chromophore-protein complexes
has a significance for their light-harvesting efficiency, and these
systems evolved to optimize it for their corresponding ecological
situation, it can be expected that the properties of at least some
parts of the photosynthetic machinery would be tuned to the fluctuation
properties of their source of excitation. For plants and some bacteria
this may be the Sun light, others like FMO complexes could be expected
to be tuned to the properties of their associated chlorosoms.
\section{Conclusions\label{sec:Conclustions}}
In this paper we have discussed dynamics of a molecular system subject
to external pumping by a light source. In particular we have considered
excitation by thermal light, by coherent pulsed light and an excitation
through a mesoscopic antenna. With a completely quantum mechanical
treatment, we have derived a general formula which enables us to study
the effect of different light properties on the photo-induced dynamics
of a molecular systems. This formula naturally contains the system--environment
interaction contribution to the excitation process which enters via
appearance of the reduced density matrix dynamics. We show that once
the properties of light are known in terms of a certain two-point
correlation function, the only information needed to reconstruct the
systems dynamics is the reduced evolution superoperator, which is
in principle accessible through ultrafast non-linear spectroscopy.
This conclusion applies to any type of light and makes thus the results
of ultrafast spectroscopic experiments universally relevant. Considering
a direct excitation of a small molecular antenna we found that excitation
of coherences is possible due to overlap of homogeneous line shapes
associated with different excitonic states. These coherences are however
static and correspond to a change of the preferred basis set into
which the system relaxes from the one defined by the bath only, to
the one defined by the action of both the light and the bath. When
an excitation of a photosynthetic complex mediated by a larger, possibly
mesoscopic, system is considered, the complex can harvest fluctuations
originating from the non-equilibrium state of the mediator. Fluctuations
of the mesoscopic system such as chlorosoms may time localize excitation
events of the energy channeling complex, and to excite adjacent energy
channeling complex coherently. It is likely that in such a case the
properties of energy channeling complexes like the well-know Fenna-Mathews-Olson
complex would be specially tuned to the fluctuation properties of
their associated chlorosoms.
\begin{acknowledgments}
This work was partially supported by the Czech Science Foundation
(GACR) through grant. nr. 206/09/0375, by the Ministry of Education,
Youth and Sports of the Czech Republic through grant KONTAKT ME899
and the research plan MSM0021620835, and by the Agency of the International
Science and Technology in Lithuania.
\bibliographystyle{prsty}
|
\section{Introduction}
The success of the Standard Model (SM) in describing the mass origin of elementary particles has satisfied many theoretical physicists but it is now challenged by the existence of neutrino oscillations observed in the solar\cite{SNO}, atmospheric\cite{SK}, reactor\cite{KM} and accelerator\cite{K2K} neutrino experiments, which provide us with convincing evidence for neutrino masses and lepton flavor mixing. The underlying nature of neutrino mixings as compared with that of the quark mixings has inspired a large amount of speculation regarding symmetries in the quark-lepton world as well as other kinds of new physics beyond the SM\cite{King:2003jb}.
In the Pontecorvo-Maki-Nakagawa-Sakata (PMNS)\cite{MNS} lepton mixing matrix, the most distinct feature is the existence of two large mixing angles, which is quite different from the pattern the Cabibbo-Kobayashi-Maskawa (CKM) \cite{CKM} quark mixing matrix. To be specific, the PMNS matrix consists of a large and nearly maximal angle $\vartheta_{23}$ (atmospheric angle), a large but non-maximal angle $\vartheta_{12}$ (solar angle), and a small angle $\vartheta_{13}$ (reactor angle) in the Standard Parametrization. An interesting phenomenological relation between the lepton and quark mixing angles, the so-called quark lepton complementarity (QLC), has been noticed recently\cite{Propose_smirnov}. Namely, the sums of the mixing angles of quarks and leptons for the 1-2 and 2-3 mixings agree with $45^\circ$:
\begin{eqnarray}
\theta_{12}+\vartheta_{12}\simeq45^\circ,~~~~~\theta_{23}+\vartheta_{23}\simeq45^\circ,
\end{eqnarray}
where $\theta_{12}$ and $\theta_{23}$ are quark mixing angles. As for the 1-3 mixing angles of quarks and leptons, a similar relation $\theta_{13}+\vartheta_{13}\simeq45^\circ$ does not hold because their sum is
less than $10^\circ$.
Attempts to understand the deep meaning behind the QLC relations have been made. It has been interpreted as an evidence for certain quark-lepton symmetry or quark-lepton unification\cite{Propose_Raidal}. Some other aspects of the QLC relations have also been discussed, such as their phenomenological implications\cite{phenomenology} and renormalization group (RG) effects\cite{Smirnov_RGE}. There are also the extended QLC relations proposed and discussed in the Seesaw Mechanisms\cite{extended_QLC}. Recent reviews about the QLC relations can be found in Ref.\cite{Review}. However, whether this relation is accident or not remains an open question. In Ref.\cite{Jarlskog} Jarlskog points out that the QLC relations are parametrization-variant and the specific models are far from being sufficiently pinned-down to be useful for connecting quark and lepton mixing angles like this.
In this paper, we intend to analyze the parametrization dependence of the QLC relations by calculating the mixing angles of each possible parametrization. Among nine angle-phase parametrizations of the CKM and PMNS matrices, we find that five of them can have the approximate QLC relations. If the QLC relations are assumed to exactly hold in a certain parametrization such as the Standard Parametrization, we examine whether they are possible to exactly hold in other parametrizations. Furthermore, the stability of the QLC relations under the RG running is also studied in the Fritzsch-Xing (FX) Parametrization\cite{FXpara}.
The remaining part of this paper is organized as follows. In Section~\ref{section:check}, with the latest experimental data for the CKM and PMNS matrices, we calculate the mixing angles and sum them up in each parametrization to check the QLC relations. Section~\ref{section:transformation} is devoted to examining the relationships between different parametrizations, especially whether the QLC relations in one parametrization can also hold in other parametrizations, and what conditions should be satisfied. The RG running effects on the QLC relations are discussed in the FX Parametrization both in the SM and the Minimal Supersymmetric Standard Model (MSSM) in Section~\ref{section:RGE}.
\section{QLC relations in different angle-phase Parametrizations}\label{section:check}
In this section, we try to make numerical calculations of the mixing angles of quark and lepton flavor mixing matrices and examine the QLC relations for all the possible angle-phase parametrizations.
The 3$\times$ 3 CKM quark mixing matrix can be expressed in terms of four independent parameters, which are usually taken as three rotation angles and one CP-violating phase angle. For a clear classification of this kind of angle-phase parametrizations, see \cite{Fritzsch-Xing}. It is pointed out that the CKM matrix $V$, if real and orthogonal, can in general be written as a product of three matrices $R_{12}$, $R_{23}$ and $R_{31}$, which describe simple rotations in the (1, 2), (2, 3) and (3, 1) planes.
\begin{eqnarray}
R_{12}(\theta)=\left(
\begin{array}{ccc}
c_\theta & s_\theta & 0 \\
-s_\theta & c_\theta & 0 \\
0 & 0 & 1
\end{array}\right),
~~R_{23}(\sigma)=\left(
\begin{array}{ccc}
1 & 0 & 0 \\
0 & c_{\sigma} & s_{\sigma} \\
0 & -s_{\sigma} & c_{\sigma}
\end{array}\right),
~~R_{31}(\tau)=\left(
\begin{array}{ccc}
c_\tau & 0 & s_\tau \\
0 & 1 & 0 \\
-s_\tau & 0 & c_\tau
\end{array}\right),
\end{eqnarray}
where $s_\theta\equiv{\rm sin}\theta,c_\theta\equiv{\rm cos}\theta$, etc. After introducing the CP-violating phase $\phi$, among all the twelve possible products only nine of them are structurally different, as the remaining three products are correlated with each other and leading essentially to the same form. And the specific forms of the nine possible angle-phase parametrizations are listed in the left column of Table~\ref{9-Parameterizations} as P1-P9, and generally P1 corresponds to the Standard Parametrization \cite{PDG} and P2 to the FX Parametrization. For Majorana neutrinos, two additional parameters are needed in PMNS lepton mixing matrix, namely two Majorana CP-violating phase angles, which do not affect oscillations\cite{Strumia}.
In the calculation of quark mixing angles, we take the Wolfenstein parametrization with the accuracy of $O(\lambda^{6})$ as proposed in \cite{Wolfenstein} which is shown as below:
\begin{equation}
V_{\rm CKM} = \left (
\begin{array}{ccc}
1-\frac{1}{2}\lambda^{2}-\frac{1}{8}\lambda^{4} & \lambda & A\lambda^{3}(\rho
-i\eta) \\
-\lambda \left [1+ \frac{1}{2}A^{2}\lambda^{4}(2\rho -1) + i
A^{2}\lambda^{4}\eta \right ] &
1-\frac{1}{2}\lambda^{2}-\frac{1}{8}\left (4A^{2}+1\right)\lambda^{4} &
A\lambda^{2} \\
A\lambda^{3}(1-\rho-i\eta) & -A\lambda^{2}\left [1+\frac{1}{2}\lambda^{2}(2\rho
-1 )
+i\lambda^{2}\eta \right ] & 1-\frac{1}{2}A^{2}\lambda^{4}
\end{array}
\right ) \;
\end{equation}
To calculate the moduli of the mixing matrix elements, we adopt the following inputs given by the Particle Data Group \cite{PDG}:
\begin{eqnarray}
\lambda = 0.2257^{+0.0009}_{-0.0010},~~~~A=0.814^{+0.021}_{-0.022},~~~~\bar{\rho}=0.135^{+0.031}_{-0.016},~~~~~~\bar{\eta}=0.349^{+0.015}_{-0.017}
\end{eqnarray}
where
\begin{eqnarray}
\nonumber
&\bar{\rho}&=\rho - \frac{1}{2}\rho \lambda^2 +\left(\frac{1}{2}A^2\rho -\frac{1}{8} \rho-A^2(\rho^2-\eta^2)\right)\lambda^4+O(\lambda^6), \\
&\bar{\eta} &=\eta-\frac{1}{2}\eta \lambda^2+\left(\frac{1}{2}A^2\eta -\frac{1}{8}\eta-2 A^2\rho\eta\right)\lambda^4+O(\lambda^6).
\end{eqnarray}\begin{tiny}\end{tiny}
Then we obtain
\begin{eqnarray}
|V_{\rm CKM}|=\left(
\begin{array}{ccc}
0.974205^{-0.00021}_{+0.00023} & 0.225700^{+0.00090}_{-0.00100} & 0.003592^{+0.00040}_{-0.00034} \\
0.225560^{+0.00090}_{-0.00100} & 0.973346^{-0.00027}_{+0.00029} & 0.041466^{+0.00141}_{-0.00148} \\
0.008733^{+0.00011}_{-0.00027} & 0.040709^{+0.00144}_{-0.00148} & 0.999140^{-0.00006}_{+0.00006}
\end{array}\right).
\end{eqnarray}
This result allows us to calculate the mixing angles in all the nine parameterizations according to the relations between angles and moduli.
For lepton mixing angles, the Standard Parameterization is expressed in terms of three mixing angles $\vartheta_{12}$, $\vartheta_{13}$, $\vartheta_{23}$ and one CP-violating phase angle $\varphi$. As shown below, the first row and third column have a pretty simple form.
\begin{eqnarray}
V_{\rm PMNS} = \left (
\begin{array}{ccc}
c^{~}_{12} c_{13} & s^{~}_{12} c_{13} & s_{13} \cr
-c^{~}_{12} s_{23} s_{13} - s^{~}_{12} c_{23} e^{-{\rm i}\varphi}
& -s^{~}_{12} s_{23} s_{13} + c^{~}_{12} c_{23} e^{-{\rm i}\varphi} & s_{23} c_{13} \cr
-c^{~}_{12} c_{23} s_{13} + s^{~}_{12} s_{23} e^{-{\rm i}\varphi}
& -s^{~}_{12} c_{23} s_{13} - c^{~}_{12} s_{23} e^{-{\rm i}\varphi} & c_{23} c_{13}
\end{array}
\right),
\end{eqnarray}
where $s_{ij}={\rm sin}\vartheta_{ij}, c_{ij}={\rm cos}\vartheta_{ij}~(i,j=1,2,3)$. With the latest global fit of the experimental data given in\cite{Fogli:2009zza}, the three mixing angles read
\begin{eqnarray}
\nonumber
{\rm sin}^2\vartheta_{12}=0.312(1^{+0.128}_{-0.109})~(2\sigma)\\
\nonumber
{\rm sin}^2\vartheta_{23}=0.466(1^{+0.292}_{-0.215}) ~(2\sigma) \\
{\rm sin}^2\vartheta_{13}=0.016\pm 0.010 ~(1\sigma)
\end{eqnarray}
Due to the smallness of $\vartheta_{13}\simeq~(7.27^{+2.012}_{-2.824})^\circ$, those terms including ${\rm sin}\vartheta_{13}$ could be neglected in the (2, 1), (2, 2), (3, 1) and (3, 2) entries in the Standard Parameterization. Thus the moduli of the mixing matrix elements can be obtained,
\begin{equation}
\left|V_{\rm PMNS}\right| = \left (
\begin{array}{ccc}
0.822795^{-0.0283}_{+0.0244} & 0.554083^{+0.0314}_{-0.0284} & 0.126491^{+0.0348}_{-0.0490} \\
0.408176^{-0.0340}_{+0.0117} &
0.606129^{-0.0983}_{+0.0705} &
0.677159^{+0.0886}_{-0.0742} \\
0.381303^{+0.0790}_{-0.0624} & 0.566223^{+0.0584}_{-0.0523} & 0.724883^{-0.1023}_{+0.1023}
\end{array}
\right ). \;
\end{equation}
With the relations between the moduli of mixing matrix elements and mixing angles in each of the parametrization, we can get all the mixing angles.
The numerical results of quark and lepton mixing angles as well as their QLC relations are listed in the right column of Table~\ref{9-Parameterizations}. It is obvious from the table that the QLC relations approximately hold in P1, P2, P3, P4 and P5 parametrizations but suffer from large deviation in the remaining four parametrizations. Thus the QLC relations are indeed parameterization-dependent. Furthermore, the distinct feature for those parametrizations accommodating the QLC relations is that they all have a simple form in their (1, 3) entries.
\section{Conditions for the exact QLC relations Transformation}\label{section:transformation}
Since the QLC relations depend on the forms of parametrizations, the exploration of those parametrizations which ensure this relation is necessary and pressing. Based on the hypothesis that the QLC relations hold in the Standard Parametrization,
\begin{eqnarray}
\theta_{12}+\vartheta_{12}=45^\circ,~~~\theta_{23}+\vartheta_{23}=45^\circ,
\label{qlc2}
\end{eqnarray}
we aim to see under what conditions this is still the case for the corresponding angles in the other parametrizations.
Take the FX Parametrization (i.e. P2 in Table.~\ref{9-Parameterizations}) as an example, we have
\begin{eqnarray}
{\rm tan}\theta_d=\frac{|V_{td}|}{|V_{ts}|}=\frac{|c_{\theta_{12}} c_{\theta_{23}} s_{\theta_{13}} - s_{\theta_{12}} s_{\theta_{23}}e^{-i\phi}|}{|s_{\theta_{12}} c_{\theta_{23}} s_{\theta_{13}} + c_{\theta_{12}} s_{\theta_{23}}e^{-i\phi}|},~~~~
{\rm cos}\theta=|V_{tb}|=|c_{\theta_{23}} c_{\theta_{13}}|.
\label{relation}
\end{eqnarray}
We first consider $\theta_d+\vartheta_\nu$ which is corresponding to the first relation in Eq.(\ref{qlc2}):
\begin{eqnarray}
\nonumber
{\rm tan}(\theta_d+\vartheta_\nu)&=& \frac{{\rm tan}\theta_d+{\rm tan}\vartheta_\nu}{1-{\rm tan}\theta_d {\rm tan}\vartheta_\nu }\\
&=&\frac{{\cal C} {\cal A}+{\cal D}{\cal B}}{{\cal D}{\cal A}-{\cal C}{\cal B}},
\label{tan}
\end{eqnarray}
where ${\cal A}$, ${\cal B}$, ${\cal C}$ and ${\cal D}$ are defined as
\begin{eqnarray}
\nonumber
{\cal A}&=&|s_{\vartheta_{12}} c_{\vartheta_{23}} s_{\vartheta_{13}}+c_{\vartheta_{12}} s_{\vartheta_{23}}e^{-i\varphi}|,~~~{\cal B}=|c_{\vartheta_{12}} c_{\vartheta_{23}}s_{\vartheta_{13}}+s_{\vartheta_{12}}s_{\vartheta_{23}}e^{-i\varphi}|, \\
{\cal C}&=&|c_{\theta_{12}} c_{\theta_{23}} s_{\theta_{13}}-s_{\theta_{12}}s_{\theta_{23}}e^{-i\phi}|,~~~~{\cal D}=|s_{\theta_{12}}c_{\theta_{23}}s_{\theta_{13}}+c_{\theta_{12}}s_{\theta_{23}}e^{-i\phi}|.
\label{ABCD}
\end{eqnarray}
Using the QLC relation in Eq.(\ref{qlc2}), one can get ${\cal A}$ and ${\cal B}$ expressed in the form of the Standard Parametrization:
\begin{eqnarray}
\nonumber
{\cal A}&=&\frac{1}{2}|(c_{\theta_{12}}-s_{\theta_{12}})(c_{\theta_{23}}+s_{\theta_{23}})s_{\vartheta_{13}}+(c_{\theta_{12}}+s_{\theta_{12}})(c_{\theta_{23}}-s_{\theta_{23}})e^{-i\varphi}|, \\
{\cal B}&=&\frac{1}{2}|(c_{\theta_{12}}+s_{\theta_{12}})(c_{\theta_{23}}+s_{\theta_{23}})s_{\vartheta_{13}}+(c_{\theta_{12}}-s_{\theta_{12}})(c_{\theta_{23}}-s_{\theta_{23}})e^{-i\varphi}|.
\end{eqnarray}
After substituting the above expressions of ${\cal A}$ and ${\cal B}$ into Eq.({\ref{tan}}), we find that it is hard to deduce any useful conclusion from it. Furthermore, we assume that the corresponding smallest angles in the Standard Parametrization for quark and lepton mixings are vanishing, i.e. $\theta_{13}=\vartheta_{13}=0^\circ$, thus
\begin{eqnarray}
\nonumber
{\cal A}&=&\frac{1}{2}|(c_{\theta_{12}}+s_{\theta_{12}})(c_{\theta_{23}}-s_{\theta_{23}})|,~~
{\cal B}=\frac{1}{2}|(c_{\theta_{12}}-s_{\theta_{12}})(c_{\theta_{23}}-s_{\theta_{23}})|, \\
{\cal C}&=&|s_{\theta_{12}}s_{\theta_{23}}|,~~~~~~~~~~~~~~~~~~~~~~~~~~
{\cal D}=|c_{\theta_{12}}s_{\theta_{23}}|,
\end{eqnarray}
which leads to a remarkable result,
\begin{eqnarray}
\nonumber
{\rm tan}(\theta_d+\vartheta_\nu)&=&\frac{{\cal C} {\cal A}+{\cal D}{\cal B}}{{\cal D}{\cal A}-{\cal C}{\cal B}} \\
\nonumber
&=&\frac{|s_{\theta_{12}}s_{\theta_{23}}| |(c_{\theta_{12}}+s_{\theta_{12}})(c_{\theta_{23}}-s_{\theta_{23}})|+|c_{\theta_{12}}s_{\theta_{23}}||(c_{\theta_{12}}-s_{\theta_{12}})(c_{\theta_{23}}-s_{\theta_{23}})|}{|c_{\theta_{12}}s_{\theta_{23}}||(c_{\theta_{12}}+s_{\theta_{12}})(c_{\theta_{23}}-s_{\theta_{23}})|-|s_{\theta_{12}}s_{\theta_{23}}||(c_{\theta_{12}}-s_{\theta_{12}})(c_{\theta_{23}}-s_{\theta_{23}})|} \\
\nonumber
&=&\left|\frac{s_{\theta_{12}}(c_{\theta_{12}}+s_{\theta_{12}})+c_{\theta_{12}}(c_{\theta_{12}}-s_{\theta_{12}})}{c_{\theta_{12}}(c_{\theta_{12}}+s_{\theta_{12}})-s_{\theta_{12}}(c_{\theta_{12}}-s_{\theta_{12}})}\right| \\
&=&1,
\end{eqnarray}
from which the QLC relation in the FX Parametrization $\theta_d+\vartheta_\nu=45^\circ$ exactly holds.
Now we turn to consider the second relation of Eq.(\ref{relation}),
\begin{eqnarray}
\nonumber
{\rm cos}(\theta+\vartheta) &=& c_{\theta} c_{\vartheta} -s_{\theta} s_{\vartheta}\\
\nonumber
&=&c_{\theta_{23}}c_{\theta_{13}}c_{\vartheta_{23}}c_{\vartheta_{13}}-\sqrt{(1-c_{\theta_{23}}^2c_{\theta_{13}}^2)(1-c_{\vartheta_{23}}^2c_{\vartheta_{13}}^2)} \\
\nonumber
&=&\frac{\sqrt{2}}{2}c_{\theta_{23}}c_{\theta_{13}}c_{\vartheta_{13}}(c_{\theta_{23}}+s_{\theta_{23}})-\sqrt{(1-c_{\theta_{23}}^2c_{\theta_{13}}^2)\left[1-\frac{1}{2}(c_{\theta_{23}}+s_{\theta_{23}})^2c_{\vartheta_{13}}^2\right]}.
\end{eqnarray}
With the help of the QLC relation for $\theta_{23}$ and $\vartheta_{23}$ and the assumption of the vanishing smallest mixing angles $\theta_{13}=\vartheta_{13}=0^\circ$, one can obtain
\begin{eqnarray}
\nonumber
{\rm cos}(\theta+\vartheta) &=& \frac{\sqrt{2}}{2}c_{\theta_{23}}(c_{\theta_{23}}+s_{\theta_{23}})-\sqrt{s_{\theta_{23}}^2\left[1-\frac{1}{2}(1-2c_{\theta_{23}}s_{\theta_{23}})\right]} \\
\nonumber
&=&\frac{\sqrt{2}}{2}c_{\theta_{23}}(c_{\theta_{23}}+s_{\theta_{23}})-\frac{\sqrt{2}}{2}s_{\theta_{23}}(c_{\theta_{23}}-s_{\theta_{23}})\\
\nonumber
&=&\frac{\sqrt{2}}{2}.
\end{eqnarray}
Again, the QLC relation holds for the FX Parametrization in this situation. Namely,
\begin{eqnarray}
\theta_d+\vartheta_\nu=45^\circ~~ {\rm and}~~\theta+\vartheta=45^\circ.
\end{eqnarray}
In fact, under the condition of $\theta_{13}=\vartheta_{13}=0^\circ$, the conclusion that the QLC relations hold in these parametrizatons can be exactly obtained. For example, from Eq.(\ref{relation}) we can easily get ${\rm tan}\theta_d=|{\rm tan}\theta_{12}|$ and ${\rm cos}\theta=|{\rm cos}\theta_{23}|$ in the FX Parametrization if $\theta_{13}=\vartheta_{13}=0^\circ$. So is the case in the lepton sector, i.e. ${\rm tan}\vartheta_\nu=|{\rm tan}\vartheta_{12}|$ and ${\rm cos}\vartheta=|{\rm cos}\vartheta_{23}|$. Hence, the QLC relations hold in the FX Parametrization. And the same conclusion can also be obtained for P3, P4 and P5 Parametrizations in a similar procedure. The reason is simple that these five parametrizations are essentially equivalent to one another in the $\theta_{13}=\vartheta_{13}=0^\circ$ limit.
\section{on the stability of QLC relations RG running}\label{section:RGE}
As proposed in many papers, the quark-lepton symmetry implied by the QLC relations means that physics responsible for these relations should be realized at some scales which might be the quark-lepton unification scale, $\Lambda_{\rm GUT}$ or even higher scales, and the RG effects has been discussed in the framework of the Standard Parametrization\cite{Smirnov_RGE,Antusch:2003kp}. Since there are specific advantages in the FX Parametrization for the study of fermion mass matrices and B-meson physics\cite{Fritzsch-Xing}, it is useful to examine the sensitivity of the QLC relations to the RG effects in this parametrization. And it has been shown that the RG equations of quark and lepton mixing angles have a particularly simple form in the FX Parametrization\cite{XRGE_q,XRGE_l}. Assume that QLC relations hold exactly at the scale $M_Z$ in this parametrization:
\begin{eqnarray}
\theta_d+\vartheta_\nu = 45^\circ,~~~\theta+\vartheta =45^\circ,
\label{assumption_3}
\end{eqnarray}
thus
\begin{eqnarray}
\dot{\theta}_d+\dot{\vartheta}_\nu=0,~~~\dot{\theta}+\dot{\vartheta}=0,
\label{qlc_RGE2}
\end{eqnarray}
where $\dot{\theta}=\displaystyle{\frac{{\rm d}\theta}{{\rm d}t}}$ with $t\equiv {\rm ln}~(\mu/M_Z)$. We already have the RG equations of three quark mixing angles \cite{XRGE_q} and three Dirac neutrino mixing angles \cite{XRGE_l} in FX Parametrization:
\begin{eqnarray}
\renewcommand\arraystretch{1.5}
\left\{\begin{array}{l}
\dot{\theta}_u =
\displaystyle{-\frac{1}{32\pi^2} }C y^2_b \sin 2\theta_u \sin^2\theta, \\
\dot{\theta}_d = \displaystyle{-\frac{1}{32\pi^2 }} C y^2_t \sin 2\theta_d \sin^2\theta, \\
\dot{\theta} = \displaystyle{-\frac{1}{32\pi^2}} C
\left( y^2_b + y^2_t \right) \sin 2\theta,
\end{array}\right.~~~~~~~~
\left\{\begin{array}{l}
\dot{\vartheta}_l = \displaystyle{+ \frac{Cy^2_\tau}{16\pi^2}} ~ c^{}_{\nu}
s^{}_{\nu} c_{\vartheta} c^{}_{\varphi} \left ( \xi^{}_{13} - \xi^{}_{23} \right ), \\
\dot{\vartheta}_\nu = \displaystyle{ + \frac{Cy^2_\tau}{16\pi^2}} ~ c^{}_{\nu}
s^{}_{\nu} \left [ s_{\vartheta}^2 \xi^{}_{12} + c_{\vartheta}^2 \left ( \xi^{}_{13} -
\xi^{}_{23} \right ) \right ], \\
\dot{\vartheta} = \displaystyle{ + \frac{Cy^2_\tau}{16\pi^2}} ~ c_{\vartheta} s_{\vartheta} \left (
s^2_{\nu} \xi^{}_{13} + c^2_{\nu} \xi^{}_{23} \right ),
\end{array}\right.
\label{angleRGE}
\end{eqnarray}
where $C=-1.5~(+1)$ in the SM (MSSM), $\xi^{}_{ij} \equiv \left (y^2_i + y^2_j \right )/\left
(y^2_i - y^2_j \right )$ and $y_\alpha,~y_a $ and $y_i$ $(\alpha=\tau,a=b,t$ and $ i=1,2,3)$ stand respectively for the eigenvalues of the Yukawa coupling matrices of charged leptons, quarks and neutrinos. In the case of the SM, the Yukawa couplings $y_i=\displaystyle{\frac{m_i}{v}}(i=1,2,3)$, where the Higgs vacuum expection value (VEV) $v$ is $174$ GeV. In the MSSM, $m_\alpha=y_\alpha v {\rm sin}\beta,~m_\gamma=y_{\gamma} v {\rm cos}\beta~(\alpha=u,c,t,\gamma=d,s,b,e,\mu,\tau)$, where tan$\beta$ is the ratio of two Higgs VEV's.
Some qualitative comments on the main features of Eq.(\ref{angleRGE}) are in order.
(a) For the RG equations of quark flavor mixing angles in both the SM and MSSM, noticing that the value of $\theta$ is very small, we can safely claim that the RG running effects of $\theta_u$, $\theta_d$ and $\theta$ are highly suppressed. As a result, three quark mixing angles in FX Parametrization will not change a lot under the RG running.
(b) In the lepton sector in the SM case, the derivatives of three mixing angles are proportional to $y_\tau^2=\displaystyle{\left(\frac{m_\tau}{v}\right)^2}\simeq10^{-4}$\cite{Xing_running}. Notice that $\xi_{ij}=-\displaystyle{\frac{m_i^2+m_j^2}{\Delta m_{ji}^2}}$, with $\Delta m_{ji}^2=m_j^2-m_i^2$, $\Delta m_{21}^2\simeq7.7\times10^{-5}~{\rm eV^2}$ and $|\Delta m_{32}^2|\approx|\Delta m_{31}^2|\approx 2.4\times 10^{-3}~{\rm eV^2} $\cite{Fogli:2009zza}. Thus the most sensitive angle to radiative corrections is $\vartheta_\nu$ whose RG equation is the only one that consists of $\xi_{12}$. But we still cannot expect large running effects on $\theta_{\nu}$ because the loop factor $1/16\pi^2$ makes the derivative even smaller. While in the MSSM case, where $y_\tau=\displaystyle{\frac{m_\tau}{v {\rm cos}\beta}}$, the RG running effects could be enhanced when ${\rm tan}\beta$ is significantly large.
As a result, if we sum up the derivatives of the corresponding mixing angles of quark and lepton sectors in Eq.(\ref{angleRGE}), we can conclude that in the SM case, the QLC relations are essentially stable under the RG running and in the MSSM case, these relations might become unstable only when ${\rm tan}\beta$ is sufficiently large.
\section{summary}\label{section:summary}
To understand the deep meaning of the quark and lepton world, the quark-lepton symmetry topic has drawn a lot of attention recent years. Among many of the aspects that imply the symmetry and unification in quark and lepton sectors, the QLC relations between the mixing angles of the CKM and PMNS matrices have been considered very interesting and suggestive.
In this paper, we have calculated the QLC relations for each of the angle-phase parametrizations and find that these relations are parametrization-dependent. Furthermore, the distinct feature of those parametrizations which can approximately accommodate the QLC relations is that they all have a simple form in the (1, 3) entries. Then based on the assumption that the QLC relations hold exactly in the Standard Parametrization we make an exploration in the FX Parametrization and get the conclusion that these relations can also hold as long as the smallest mixing angle $\theta_{13}$ is vanishing. Finally, we make clear that the QLC relations can essentially stay stable under the RG running effects in the SM and MSSM unless the value of ${\rm tan}\beta$ is sufficiently large.
\begin{acknowledgements}
The author is greatly indebted to Professor Z.Z. Xing for stimulating discussions, constant instruction and polishing the manuscript with great care, and to S. Luo for warm hospitality during her visiting stay in Beijing IHEP where this work was done. She is also grateful to W.T. Deng, Y.K. Song, H. Zhang and Professor Z.G. Si for useful discussions and comments. This work is supported in part by NSFC and Natural Science Foundation of Shandong Province.
\end{acknowledgements}
\small
\begin{table}\label{table}
\caption{Classification of different parametrizations for the flavor mixing
matrix and the QLC relations.}
\vspace{-1cm}
\begin{center}
\begin{tabular}{ccc} \\ \hline\hline
Parametrization &~& Quark Lepton Complementarity
\\ \hline
\underline{{\it P1:} ~ $V \; = \; R_{23}(\theta_{23}) ~ R_{31}(\theta_{13}, \phi)
~ R_{12}(\theta_{12})$}
&& $\theta_{12}/\theta_{23}/\theta_{13}~~~~~~~~\vartheta_{12}/\vartheta_{23}/\vartheta_{13}~~~~~~~~~~~~~~~~~~~~~~$
\\
$\left ( \matrix{
c^{~}_{12} c_{13} & s^{~}_{12} c_{13} & s_{13} \cr
-c^{~}_{12} s_{23} s_{13} - s^{~}_{12} c_{23} e^{-{\rm i}\phi}
& -s^{~}_{12} s_{23} s_{13} + c^{~}_{12} c_{23} e^{-{\rm i}\phi} & s_{23} c_{13} \cr
-c^{~}_{12} c_{23} s_{13} + s^{~}_{12} s_{23} e^{-{\rm i}\phi}
& -s^{~}_{12} c_{23} s_{13} - c^{~}_{12} s_{23} e^{-{\rm i}\phi} & c_{23} c_{13} \cr} \right )
$
&& $\matrix{
(13.04^{+0.053}_{-0.059})^\circ+ (33.96^{+2.430}_{-2.137})^\circ=(47.00^{+2.483}_{-2.196})^\circ \cr
(2.37^{+0.081}_{-0.085})^\circ+(43.05^{+7.839}_{-5.834})^\circ=(45.42^{+7.920}_{-5.919})^\circ \cr
(0.20^{+0.023}_{-0.020})^\circ+(7.27^{+2.012}_{-2.824})^\circ=(7.47^{+2.035}_{-2.844})^\circ \cr} $ \\ \\
\underline{{\it P2:} ~ $V \; = \; R_{12}(\theta_u) ~ R_{23}( \theta, \phi)
~ R^{-1}_{12}(\theta_d)$}
&&$\theta_u/\theta_d/\theta~~~~~~~~~~\vartheta_l/\vartheta_\nu/\vartheta~~~~~~~~~~~~~~~~~~~~~~ $\\
$\left ( \matrix{
s^{~}_{u} s^{~}_{d} c_{ } + c^{~}_{u} c^{~}_{d} e^{-{\rm i}\phi} &
s^{~}_{u} c^{~}_{d} c_{ } - c^{~}_{u}
s^{~}_{d} e^{-{\rm i}\phi} & s^{~}_{u} s_{ } \cr
c^{~}_{u} s^{~}_{d} c_{ } - s^{~}_{u} c^{~}_{d} e^{-{\rm i}\phi} &
c^{~}_{u} c^{~}_{d} c_{ } + s^{~}_{u}
s^{~}_{d} e^{-{\rm i}\phi} & c^{~}_{u} s_{ } \cr
- s^{~}_{d} s_{ } & - c^{~}_{d}
s_{ } & c_{ } \cr} \right )
$
&& $\matrix{
(4.95^{+0.363}_{-0.305})^\circ+(10.58^{+1.310}_{-3.261})^\circ=(15.53^{+1.637}_{-3.566})^\circ \cr
(12.11^{-0.262}_{+0.065})^\circ+(33.96^{+2.430}_{-2.137})^\circ=(46.67^{+2.168}_{-2.072})^\circ \cr
(2.38^{+0.081}_{-0.085})^\circ+(43.54^{+7.956}_{-6.099})^\circ=(45.92^{+8.037}_{-6.184})^\circ \cr}$ \\ \\
\underline{{\it P3:} ~ $V \; = \; R_{23}(\theta_d) ~ R_{12}(\theta, \phi)
~ R^{-1}_{23}(\theta_u)$}
&& $\theta_u/\theta_d/\theta~~~~~~~~~~\vartheta_l/\vartheta_\nu/\vartheta~~~~~~~~~~~~~~~~~~~~~~ $ \\
$\left ( \matrix{
c^{~}_{\theta} & s^{~}_{\theta} c_{u} & -s^{~}_{\theta} s_{u} \cr
-s^{~}_{\theta} c_{d} & c^{~}_{\theta} c_{d} c_{u} + s_{d} s_{u} e^{-{\rm i}\phi}
& -c^{~}_{\theta} c_{d} s_{u} + s_{d} c_{u} e^{-{\rm i}\phi} \cr
s^{~}_{\theta} s_{d} & -c^{~}_{\theta} s_{d} c_{u} + c_{d} s_{u} e^{-{\rm i}\phi}
& c^{~}_{\theta} s_{d} s_{u} + c_{d} c_{u} e^{-{\rm i}\phi} \cr} \right )
$
&& $\matrix{
(0.91^{+0.096}_{-0.083})^\circ+(12.86^{+2.538}_{-4.477})^\circ=(13.77^{+2.634}_{-4.560})^\circ \cr
(2.22^{+0.019}_{-0.059})^\circ+(43.05^{+7.839}_{-5.834})^\circ=(45.27^{+7.858}_{-5.893})^\circ\cr
(13.04^{+0.053}_{-0.059})^\circ+(34.63^{+2.758}_{-2.538})^\circ=(47.67^{+2.811}_{-2.597})^\circ \cr} $ \\ \\
\underline{{\it P4:} ~ $V \; = \; R_{23}(\sigma) ~ R_{12}(\theta, \phi)
~ R^{-1}_{31}(\rho)$}
&& $\theta_u/\theta_d/\theta~~~~~~~~~~\vartheta_l/\vartheta_\nu/\vartheta~~~~~~~~~~~~~~~~~~~~~~ $
\\
$\left ( \matrix{
c^{~}_{\theta} c_{u} & s^{~}_{\theta} & -c^{~}_{\theta} s_{u} \cr
-s^{~}_{\theta} c_{d} c_{u} + s_{d} s_{u} e^{-{\rm i}\phi} & c^{~}_{\theta} c_{d}
& s^{~}_{\theta} c_{d} s_{u} + s_{d} c_{u} e^{-{\rm i}\phi} \cr
s^{~}_{\theta} s_{d} c_{u} + c_{d} s_{u} e^{-{\rm i}\phi} & -c^{~}_{\theta} s_{d}
& -s^{~}_{\theta} s_{d} s_{u} + c_{d} c_{u} e^{-{\rm i}\phi} \cr} \right )
$
&& $\matrix{
(0.21^{+0.023}_{-0.020})^\circ+(8.74^{+2.733}_{-3.516})^\circ=(8.95^{+2.756}_{-3.536})^\circ\cr
(2.39^{+0.086}_{-0.088})^\circ+(43.05^{+7.839}_{-5.834})^\circ=(45.44^{+7.925}_{-5.922})^\circ \cr
(13.04^{+0.053}_{-0.059})^\circ+(33.65^{+2.189}_{-1.935})^\circ=(46.69^{+2.242}_{-1.994})^\circ \cr} $ \\ \\
\underline{{\it P5:} ~ $V \; = \; R_{31}(\rho) ~ R_{23}(\sigma, \phi)
~ R^{-1}_{12}(\theta)$}
&& $\theta_u/\theta_d/\theta~~~~~~~~~~\vartheta_l/\vartheta_\nu/\vartheta~~~~~~~~~~~~~~~~~~~~~~ $
\\
$\left ( \matrix{
-s^{~}_{\theta} s_{d} s_{u} + c^{~}_{\theta} c_{u} e^{-{\rm i}\phi}
& -c^{~}_{\theta} s_{d} s_{u} - s^{~}_{\theta} c_{u} e^{-{\rm i}\phi} & c_{d} s_{u} \cr
s^{~}_{\theta} c_{d} & c^{~}_{\theta} c_{d} & s_{d} \cr
-s^{~}_{\theta} s_{d} c_{u} - c^{~}_{\theta} s_{u} e^{-{\rm i}\phi}
& -c^{~}_{\theta} s_{d} c_{u} + s^{~}_{\theta} s_{u} e^{-{\rm i}\phi} & c_{d} c_{u} \cr} \right )
$
&& $\matrix{
(0.21^{+0.023}_{-0.020})^\circ+(9.90^{+4.622}_{-4.326})^\circ=(10.11^{+4.645}_{-4.346})^\circ \cr
(2.38^{+0.081}_{-0.085})^\circ+(42.62^{+7.354}_{-5.537})^\circ=(45.00^{+7.435}_{-5.622})^\circ \cr
(13.05^{+0.054}_{-0.059})^\circ+(33.96^{+2.430}_{-2.137})^\circ=(47.01^{+2.484}_{-2.196})^\circ \cr} $ \\ \\
\underline{{\it P6:} ~ $V \; = \; R_{12}(\theta) ~ R_{31}(\theta_u, \phi)
~ R^{-1}_{23}(\theta_d)$}
&& $\theta_u/\theta_d/\theta~~~~~~~~~~\vartheta_l/\vartheta_\nu/\vartheta~~~~~~~~~~~~~~~~~~~~~~ $
\\
$\left ( \matrix{
c^{~}_{\theta} c_{u} & c^{~}_{\theta} s_{d} s_{u} + s^{~}_{\theta} c_{d} e^{-{\rm i}\phi}
& c^{~}_{\theta} c_{d} s_{u} - s^{~}_{\theta} s_{d} e^{-{\rm i}\phi} \cr
-s^{~}_{\theta} c_{u} & -s^{~}_{\theta} s_{d} s_{u} + c^{~}_{\theta} c_{d} e^{-{\rm i}\phi}
& -s^{~}_{\theta} c_{d} s_{u} - c^{~}_{\theta} s_{d} e^{-{\rm i}\phi} \cr
-s_{u} & s_{d} c_{u} & c_{d} c_{u} \cr} \right )
$
&& $\matrix{
(0.50^{+0.006}_{-0.015})^\circ+(22.41^{+4.993}_{-3.818})=(22.91^{+4.999}_{-3.833})^\circ \cr
(2.33^{+0.083}_{-0.085})^\circ+(37.99^{+7.102}_{-5.080})^\circ=(40.32^{+7.185}_{-5.165})^\circ \cr
(13.04^{+0.053}_{-0.059})^\circ+(26.39^{-1.164}_{-0.021})^\circ=(39.43^{-1.111}_{-0.080})^\circ \cr} $ \\ \\
\underline{{\it P7:} ~ $V \; = \; R_{31}(\theta_u) ~ R_{12}(\theta, \phi)
~ R^{-1}_{31}(\theta_d)$}
&& $\theta_u/\theta_d/\theta~~~~~~~~~~\vartheta_l/\vartheta_\nu/\vartheta~~~~~~~~~~~~~~~~~~~~~~ $
\\
$\left ( \matrix{
c^{~}_{\theta} c_{u} c_{d} + s_{u} s_{d} e^{-{\rm i}\phi} & s^{~}_{\theta} c_{u}
& -c^{~}_{\theta} c_{u} s_{d} + s_{u} c_{d} e^{-{\rm i}\phi} \cr
-s^{~}_{\theta} c_{d} & c^{~}_{\theta} & s^{~}_{\theta} s_{d} \cr
-c^{~}_{\theta} s_{u} c_{d} + c_{u} s_{d} e^{-{\rm i}\phi} & -s^{~}_{\theta} s_{u}
& c^{~}_{\theta} s_{u} s_{d} + c_{u} c_{d} e^{-{\rm i}\phi} \cr} \right )
$
&& $\matrix{
(10.22^{+0.315}_{-0.320})^\circ+(45.62^{+1.233}_{-1.268})^\circ=(55.84^{+1.537}_{-1.577})^\circ \cr
(10.42^{+0.304}_{-0.320})^\circ+(58.92^{+5.037}_{-3.769})^\circ=(69.34^{+5.341}_{-4.089})^\circ \cr
(13.26^{+0.067}_{-0.073})^\circ+(52.69^{+6.791}_{-5.274})^\circ=(65.95^{+6.858}_{-5.347})^\circ \cr} $ \\ \\
\underline{{\it P8:} ~ $V \; = \; R_{12}(\theta) ~ R_{23}(\theta_d, \phi)
~ R_{31}(\theta_u)$}
&& $\theta_u/\theta_d/\theta~~~~~~~~~~\vartheta_l/\vartheta_\nu/\vartheta~~~~~~~~~~~~~~~~~~~~~~ $
\\
$\left ( \matrix{
-s^{~}_{\theta} s_{d} s_{u} + c^{~}_{\theta} c_{u} e^{-{\rm i}\phi} & s^{~}_{\theta} c_{d} &
s^{~}_{\theta} s_{d} c_{u} + c^{~}_{\theta} s_{u} e^{-{\rm i}\phi} \cr
-c^{~}_{\theta} s_{d} s_{u} - s^{~}_{\theta} c_{u} e^{-{\rm i}\phi} & c^{~}_{\theta} c_{d} &
c^{~}_{\theta} s_{d} c_{u} - s^{~}_{\theta} s_{u} e^{-{\rm i}\phi} \cr
-c_{d} s_{u} & -s_{d} & c_{d} c_{u} \cr} \right )
$
&& $\matrix{
(0.50^{+0.006}_{-0.016})^\circ+(27.75^{+8.734}_{-5.863})^\circ=(28.25^{+8.740}_{-5.879})^\circ \cr
(2.33^{+0.083}_{-0.085})^\circ+(34.49^{+4.169}_{-3.562})^\circ=(36.82^{+4.252}_{-3.647})^\circ \cr
(13.06^{+0.054}_{-0.060})^\circ+(42.43^{+6.631}_{-4.590})^\circ=(55.49^{+6.685}_{-4.650})^\circ \cr} $ \\ \\
\underline{{\it P9:} ~ $V \; = \; R_{31}(\rho) ~ R_{12}(\theta, \phi)
~ R_{23}(\sigma)$}
&& $\theta_u/\theta_d/\theta~~~~~~~~~~\vartheta_l/\vartheta_\nu/\vartheta~~~~~~~~~~~~~~~~~~~~~~ $
\\
$\left ( \matrix{
c^{~}_{\theta} c_{u} & s^{~}_{\theta} c_{\sigma} c_{u} - s_{\sigma} s_{u} e^{-{\rm i}\phi}
& s^{~}_{\theta} s_{\sigma} c_{u} + c_{\sigma} s_{u} e^{-{\rm i}\phi} \cr
-s^{~}_{\theta} & c^{~}_{\theta} c_{\sigma} & c^{~}_{\theta} s_{\sigma} \cr
-c^{~}_{\theta} s_{u} & -s^{~}_{\theta} c_{\sigma} s_{u} - s_{\sigma} c_{u} e^{-{\rm i}\phi}
& -s^{~}_{\theta} s_{\sigma} s_{u} + c_{\sigma} c_{u} e^{-{\rm i}\phi} \cr} \right )
$
&& $\matrix{
(0.51^{+0.007}_{-0.016})^\circ+(24.86^{+5.223}_{-4.236})^\circ=(25.37^{+5.230}_{-4.252})^\circ \cr
(2.43^{+0.084}_{-0.088})^\circ+(48.17^{+8.282}_{-6.463})^\circ=(50.60^{+8.366}_{-6.551})^\circ \cr
(13.04^{+0.053}_{-0.059})^\circ+(24.09^{-2.114}_{+0.737})^\circ=(37.13^{-2.061}_{+0.678})^\circ \cr} $ \\ \\
\hline\hline
\end{tabular}
\end{center}
\label{9-Parameterizations}
\end{table}
\normalsize
|
\section{Introduction}
An $n$-leaf Cladogram is an unrooted tree with $n\ge 4$ labeled leaves (vertices with degree one) and $(n-2)$ other unlabeled vertices (internal branchpoints) of degree three (see below for a figure). The number of edges in such a tree is exactly $2n-3$. Sometimes they are also referred to as phylogenetic trees. Aldous in \cite{A00} proposes the following model of a reversible Markov chain on the space of all $n$-leaf Cladograms which consists of removing a random leaf (and its incident edge) and reattaching it to one of the remaining random edges.
For a precise description we first define two operations on Cladograms. More details, with figures, can be found in \cite{A00}.
\begin{enumerate}
\item[(i)] To \textit{remove a leaf} $i$. The leaf $i$ is attached by an edge $e_1$ to a branchpoint $b$ where two other edges $e_2$ and $e_3$ are incident. Delete edge $e_1$ and branchpoint $b$, and then merge the two remaining edges $e_2$ and $e_3$ into a single edge $e$. The resulting tree has $2n-5$ edges.
\item[(ii)] To \textit{add a leaf} to an edge $f$. Create a branchpoint $b'$ which splits the edge $f$ into two edges $f_2, f_3$ and attach the leaf $i$ to branchpoint $b'$ via a new edge $f_1$. This restores the number of leaves and edges to the tree.
\end{enumerate}
Let $\mathbf{T}_n$ denote the finite collection of all $n$-leaf Cladograms. Write $\mathbf{t}' \sim \mathbf{t}$ if $\mathbf{t}'\neq \mathbf{t}$ and $\mathbf{t}'$ can be obtained from $\mathbf{t}$ by following the two operations above for some choice of $i$ and $f$. Thus a $\mathbf{T}_n$ valued chain can be described by saying: remove leaf $i$ uniformly at random, then pick edge $f$ at random and re-attach $i$ to $f$. If we assume every edge to be of unit length, then it also involves resizing the edge length after every operation. In particular the transition matrix of this Markov chain is
\[
P(\mathbf{t}, \mathbf{t}')=\begin{cases}
\frac{1}{n(2n-5)},& \quad \text{if}\quad \mathbf{t}'\sim \mathbf{t}\\
\frac{n}{n(2n-5)},& \quad \text{if}\quad \mathbf{t}'=\mathbf{t}.
\end{cases}
\]
This leads to a symmetric, aperiodic, and irreducible finite state space Markov chain. Schweinsberg \cite{S} proved that the relaxation time for this chain is $O(n^2)$, improving a previous result in \cite{A00}.
In his webpage \cite{AOP} Aldous asks the following question: what is an appropriate diffusion limit of this Markov chain? The invariant distribution for the Markov chain on $n$-leaf Cladograms is clearly the Uniform distribution. It is known, see Aldous \cite{A93}, that the sequence of Uniform distributions on $n$-leaf Cladograms converge weakly to the law of the (Brownian) Continuum Random Tree (CRT). Hence, it is natural to look for an appropriate Markov process on the support of the CRT which can be thought of as a limit of the sequence of Markov chains described above. At this point it is important to understand that the support of the CRT consists of compact real trees with a measure describing the distribution of leaves. These trees are called continuum trees. For a formal definition of these concepts we refer the reader to the seminal work by Aldous in \cite{A93}. However, for an intuitive visualization, one should think of a typical continuum tree as a compact metric space on which branchpoints are dense and all edges are infinitesimally small. This implies that the Markov process that mimics the operation of removing and inserting a new leaf on a continuum tree should not jump, in other words, we can call it a diffusion.
A detailed description of this diffusion on continuum trees is forthcoming in Pal \cite{palCT}. In this article we consider several important features of this limiting diffusion that are of interest by themselves and provide bedrock for the followup construction.
\begin{figure}[t]\label{fig_tree}
\centering
\includegraphics[width=3in, height=1.5in]{Cladogram.pdf}
\caption{A 7-leaf Cladogram}
\end{figure}
Consider the branchpoint $b$ in the $7$-leaf Cladogram $\mathbf{t}$ in Figure 1. It divides the collection of leaves naturally into three sets. Let $X(\mathbf{t})=(X_1, X_2, X_3)(\mathbf{t})$ denote the vector of proportion of leaves in each set. The corresponding number of edges in these sets are $(2nX_1-1, 2nX_2-1, 2nX_3-1)$. For example, at time zero in our given tree, going clockwise from the right we have $X(0)=(3/7, 2/7, 2/7)$.
Let $\mathbb{S}_n$ denote the unit simplex
\begin{equation}\label{whatisusimp}
\mathbb{S}_n = \left\{ x\in \mathbb{R}^n:\quad x_i \ge 0 \quad \text{for all $i$ and}\quad \sum_{i=1}^n x_i=1 \right\}.
\end{equation}
Some simple algebra will reveal that for any point $x=(x_1,x_2,x_3)$ in $\mathbb{S}_3$, given $X(\mathbf{t})=x$, the difference $X_1(\mathbf{t}')- X_1(\mathbf{t})$ can only take values in $\{ -1/n,0,1/n\}$ with corresponding probabilities
\[
q_{x_1}= x_1 \frac{2n(1-x_1)-2}{2n-5}, \quad 1-p_{x_1}-q_{x_1}, \quad p_{x_1}=(1-x_1)\frac{2n x_1-1}{2n-5}.
\]
Thus
\begin{equation}\label{cladopara}
\begin{split}
E\left( X_1(\mathbf{t}') - X_1(\mathbf{t}) \mid X(\mathbf{t})=x\right)&= \frac{1}{n}\frac{ 2x_1 - (1-x_1)}{2n-5}\approx-\frac{1}{n^2}\frac{1}{2}\left( 1-3x_1 \right)\\
E\left( ( X_1(\mathbf{t}') - X_1(\mathbf{t}))^2 \mid X(\mathbf{t})=x\right)&= \frac{1}{n^2}\frac{4nx_1(1-x_1)-x_1 - 1 }{2n-5}\\
&\approx \frac{1}{n^2} 2x_1(1-x_1).
\end{split}
\end{equation}
If we take scaled limits, as $n$ goes to infinity, of the first two conditional moments (the mixed moments can be similarly verified), it is intuitive (and follows by standard tools) that as $n$ goes to infinity, this Markov chain (run at $n^2/2$ speed) will converge to a diffusion with a generator
\begin{equation}\label{cladogen}
\frac{1}{2}\sum_{i,j=1}^3 x_i\left( 1\{i=j\} - x_j\right) \frac{\partial^2}{\partial x_i \partial x_j} - \frac{1}{2}\sum_{i=1}^n \frac{1}{2}\left(1 - 3 x_i \right)\frac{\partial}{\partial x_i}.
\end{equation}
The generator written as above is similar to the generator for the well-known diffusion limit of the Wright-Fisher (WF) Markov chain models in population genetics. The WF model is one of the most popular models in population genetics. This is a multidimensional Markov chain which keeps track of the vector of proportions of certain genetic traits in a population of non-overlapping generations. A good source for an introduction to these models is Chapter 1 in the book by Durrett \cite{durrettgenetics}. For computational purposes one often takes recourse to a diffusion approximation, which, in its standard form, leads to a family of diffusions parametrized by $n$ ``mutation rates''.
The state space of the diffusion is given by $\mathbb{S}_n$ and is parametrized by a vector $(\delta_1, \ldots, \delta_n)$ of nonnegative entries. A weak solution of the WF diffusion with parameters $\delta=(\delta_1, \ldots, \delta_n)$ solves the following stochastic differential equation for $i=1,2,\ldots,n$:
\begin{equation}\label{whatisjacobi}
dJ_i(t) = \frac{1}{2}\left(\delta_i - \delta_0 J_i(t)\right) dt + \sum_{j=1}^n \tilde{\sigma}_{i,j}(J) d\beta_j(t), \qquad \delta_0 = \sum_{i=1}^n \delta_i.
\end{equation}
Here $\beta=(\beta_1, \ldots, \beta_n)$ is a standard multidimensional Brownian motion and the diffusion matrix $\tilde{\sigma}$ is given by
\begin{equation}\label{whatistsigma}
\tilde{\sigma}_{i,j}(x)= \sqrt{x_i}\left(1\{i=j\} - \sqrt{x_ix_j} \right), \quad 1\le i,j\le n.
\end{equation}
We define the Wright-Fisher diffusion with \textit{negative mutation rates} to be a family of $n$-dimensional diffusions, parametrized by $n$ nonnegative parameters $\delta=(\delta_1, \ldots, \delta_n)$, which is a weak solution of the following differential equation:
\begin{equation}\label{whatisnwf}
d\mu_i(t) = -\frac{1}{2}\left(\delta_i - \delta_0 \mu_i(t)\right) dt + \sum_{j=1}^n \tilde{\sigma}_{i,j}(\mu) d\beta_j(t), \qquad \delta_0 = \sum_{i=1}^n \delta_i.
\end{equation}
The initial condition $\mu(0)$ is in the interior of $\mathbb{S}_n$ and the process has a drift that pushes it outside the simplex. We will show later that the process is sure to hit the boundary of the simplex at which point we stop it. In the next section we will explicitly construct a weak solution of \eqref{whatisnwf}. The uniqueness in law of such a solution until it hits the boundary follows since the drift and the diffusion coefficients are smooth (hence, Lipschitz) inside the open unit simplex. The law of this process will then be denoted uniquely by NWF$(\delta_1, \ldots, \delta_n)$.
Equivalently this process can be identified by its Markov generator. Expanding $\tilde{\sigma} \tilde{\sigma}'$ and using the fact that $\sum_{i=1}^n x_i=1$, we get
\begin{equation}\label{genwf}
\mcal{A}_n = \frac{1}{2}\sum_{i,j=1}^n x_i\left( 1\{i=j\} - x_j\right) \frac{\partial^2}{\partial x_i \partial x_j} - \sum_{i=1}^n \frac{1}{2}\left(\delta_i - \delta_0 x_i \right)\frac{\partial}{\partial x_i}.
\end{equation}
which identifies \eqref{cladogen} as the generator for NWF$(1/2,1/2,1/2)$.
\bigskip
In this text we focus on properties of NWF models as a family of diffusions on the unit simplex and explore some of their properties that are important in the context of the Markov chain model on Cladograms.
\bigskip
\noindent\textbf{Part (1).} We show that, just like Wright-Fisher diffusions (see \cite{vsm}), the NWF processes can be recovered from a far simpler class of models, the Bessel-Square (BESQ) processes with negative dimensions. A comprehensive treatment of BESQ processes can be found in the book by Revuz \& Yor \cite{RY}. This family of one dimensional diffusions is indexed by a single real parameter $\theta$ (called the dimension) and are solutions of the stochastic differential equations
\begin{equation}\label{besqintro}
Z(t)= x + 2 \int_0^t \sqrt{\abs{Z(s)}}d\beta(s) + \theta t, \qquad x \ge 0, \quad t\ge 0,
\end{equation}
where $\beta$ is a one dimensional standard Brownian motion. We denote the law of this process by $Q^\theta_x$. It can be shown that the above SDE admits a unique strong solution until it hits the origin. The classical model only admits paramater $\theta$ to be non-negative. However, an extension, introduced by G\"oing-Jaeschke \& Yor \cite{yornbesq}, allows the parameter $\theta$ to be negative. It is important to note that $Q_x^\theta$ is the diffusion limit of a Galton-Watson branching process with a $\abs{\theta}$ rate of immigration (for $\theta \ge 0$) or emigration (for $\theta<0$).
In Section \ref{sec:timechange} we show that the NWF$(\delta_1,\ldots, \delta_n)$ law, starting at $(x_1,\ldots,x_n)$, can be recovered via a stochastic time-change from a collection of $n$ independent processes with laws $Q_{x_i}^{-2\delta_i}$, $i=1,\ldots,n$, and dividing each coordinate by the total sum. For the corresponding discrete models this is usually referred to as Poissonization.
In this article we utilize this relationship to infer several properties about the NWF processes. For example, we prove that these diffusions, almost surely, hit the boundary of the simplex. We derive the explicit exit density supported on the union of the boundary walls in Theorem \ref{thm:exitnwf}.
\bigskip
\noindent\textbf{Part (2).} We also prove an interesting duality relationship between WF and NWF models. To describe the duality relationship we let the NWF continue in the lower dimensional simplex when any of the coordinates hit zero. Thus, every time a coordinate hits zero, the dimension of the process gets reduced by one, and ultimately the process is absorbed at the scalar one. Such a process can be obtained by running a WF model with appropriate parameters that initially starts with dimension one and value $1$. At independent random times, the dimension of the process increases by one, and the newly added coordinate is initialized at zero. Finally we condition on the values of the process at a chosen random time. The resulting process, backwards in time and suitably time-changed, is the original NWF model.
\bigskip
\noindent\textbf{Part (3).} The time that the NWF process takes to exit the simplex is a crucial quantity due to a reason which we describe below. We keep our exposition mostly verbal without going into too much detail since the details require considerable formalism from the theory of continuum trees and will be discussed elsewhere. In \cite{palCT} we show how \textbf{Part (1)} points towards a Poissonization of the entire Aldous Markov chain which is simpler for considering scaled limits. The Poissonized version of the Markov chain on $n$-leaf Cladograms stipulates: every existing leaf has an Exponential clock of rate $2$ attached to it which determines the instances of their deaths, and every existing edge has an independent Exponential clock of rate $1$ attached to it at which point the edge is split and a new pair of vertices (one of which is a leaf) is introduced. It is an easy verification that the rates are consistent with the BESQ limit that we claimed in \textbf{Part (1)} above. Hence, one would expect that the limit of the Poissonized chains on continuum trees, normalized to give a leaf-mass measure one, and suitably time-changed would give the conjectured Aldous diffusion. This is the strategy followed in \cite{palCT}.
Now, the Poissonized chain has some beautiful and interesting structure. Please see \cite{A93} for the details about continuum trees that we use below. A continuum tree $\mathbb{T}$ comes with its associated (infinite) length measure (analogous to the Lebesgue measure) and a leaf-mass probability measure which describes how the leaves are distributed on it. We will denote the length measure by $\mathbf{Leb}(\mathbb{T})$ and the leaf-mass probability measure by $\mu(\mathbb{T})$. Suppose we sample $n$ iid elements from $\mu(\mathbb{T})$ and draw the tree generated by them, it produces an $n$-leaf Cladogram with edge-lengths (or, a proper $n$-tree, according to \cite{A93}). Thus, by using the fact that the continuum tree is compact, one can approximate a continuum tree by a sequence of $n$-leaf Cladograms.
Now consider an $n$-leaf Cladogram for a very large $n$, and further consider $m$ internal branchpoints. For example, in Figure $1$, we have three branchpoints $\{a,b,c\}$ in a $7$-leaf Cladogram. These branchpoints generate a \textit{skeleton} subtree of the original tree and partition the leaves as \textit{internal} or \textit{external} to the skeleton. The components of the vector of external leaf masses grow as independent continuous time, binary branching, Galton-Watson branching processes with a rate of branching / dying $2$ and a rate of emigration $1$. Note that this is consistent with the diffusion limit as BESQ with $\theta=-1$. As the Markov chain (Poissonized or not) proceeds, there comes a time when one of these external leaf masses gets exhausted. When this happens, one of the internal branchpoints becomes a leaf. The distribution of every coordinate of external leaf-masses at this exit time is derived in \textbf{Part (2)}. Until this time, supported on the skeleton, new subtrees can grow and decay. We show in \cite{palCT}, the dynamics of the sizes of these subtrees on the internal part can be modeled as the age process of a chronological splitting tree. Chronological splitting trees are special kind of biological trees where an individual lives up to a certain (possibly non-Exponential) lifetime and produces children at rate one during that lifetime. Her children behave in an identical manner with an independent and identically distributed lifetime of their own. The age process refers to the point process of current ages of the existing members in the family. More details about splitting trees can be found in the article by Lambert \cite{lambertAOP}.
When one of the internal vertices gets \textit{exposed}, the above dynamics breaks down and we need to find a slightly different set of internal vertices to proceed. Hence, it is important to derive estimates of the times at which this change happens.
We provide quantitative bounds on the value of this stopping time under the special situation of symmetric choice of parameters which is the case at hand.
The article is divided as follows. Our main tool in this analysis is to establish a relationship between NWF processes and Bessel-square processes of negative dimensions, much in the spirit of Pal \cite{vsm}. This has been done in Section \ref{sec:timechange} where we also establish Theorem \ref{mainthm}. The relevant results about BESQ processes have been listed in Section \ref{sec:besq}. Most of these results are known and appropriate citations have been provided. Proofs of the rest have been done in the Appendix. Exact computations of exit density from the simplex has been done in Section \ref{sec:exitdensity}. Estimates of the exit time have been established in Section \ref{sec:exittime}.
\section{Some results about BESQ processes}\label{sec:besq}
The Bessel-Square processes of negative dimensions $-\theta$, where $\theta\ge 0$, are one dimensional diffusions which are the unique strong solution of the SDE
\begin{equation}\label{sdenbesq}
X(t) = x - \theta t + 2\int_0^t \sqrt{X(s)}d\beta(s), \quad t \le T_0,
\end{equation}
where $T_0$ is the first hitting time of zero for the process $X$ and $x$ is a positive constant. The process is absorbed at zero.
We will denote the law of this process $Q_x^{-\theta}$ just as BESQ of a positive dimension $\theta$ will be denoted by $Q^{\theta}_x$.
The following collection of results is important for us. All the proofs can be found in the article by G\"oing-Jaeschke and Yor \cite{yornbesq}.
\begin{lemma}[Time-reversal]\label{lem:trev} For any $\theta > -2$ and any $x >0$, $Q_x^{-\theta}(T_0 < \infty )=1$, while, for $\theta \ge 2$, one has $Q^\theta_x(T_0 < \infty)=0$.
Moreover the following equality holds in distribution:
\begin{equation}\label{Qtreversal}
\left( X(T_0 - u), \; u \le T_0 \right) = \left( Y(u), \; u \le L_x \right),
\end{equation}
where $Y$ has law $Q^{4+\theta}_0$ and $L_x$ is the last hitting time of $x$ for the process $Y$.
In particular,
\begin{enumerate}
\item[(i)] both $L_x$ and $T_0$ are distributed as $x/2G$, where $G$ is a Gamma random variable with parameter $(\theta/2 +1)$.
\item[(ii)] The transition probabilities $p_t^\theta(x,y)$ for $x,y >0$ satisfy the identity
\[
p_t^{-\theta}(x,y)=p_t^{4+\theta}(y,x).
\]
\end{enumerate}
\end{lemma}
The following results have been proved in the Appendix.
\begin{lemma}\label{scale}
The scale function for $Q^{-\theta}$, $\theta\ge 0$, is given by the function
\[
s(x)= x^{\theta/2+1}, \quad x\ge 0.
\]
Moreover,
\begin{enumerate}
\item[(i)] The origin is an exit boundary for the diffusion and not an entry.
\item[(ii)] The change of measure
\[
x^{-\theta/2-1}Q_x^{-\theta}\left( X(t)^{\theta/2+1}1\left( \cdot \right)\right)
\]
on the $\sigma$-algebra generated by the process up to time $t$ is consistent for various $t$ and is the law of $Q_x^{4+\theta}$. Thus, we say $Q_x^{4+\theta}$ is $Q_x^{-\theta}$ conditioned never to hit zero.
\end{enumerate}
\end{lemma}
The previous fact is the generalization of the well-known observation that Brownian motion, conditioned never to hit the origin, has the law of the three-dimensional Bessel process.
\begin{lemma}\label{logct}
Let $\{Z(t), \; t\ge 0\}$ denote a BESQ process of dimension $\theta$ for some $\theta > 2$. Then
\[
\lim_{\epsilon \rightarrow 0}\frac{1}{\log(1/\epsilon)} \int_{\epsilon}^t \frac{du}{Z(u)}=\frac{1}{\theta-2}, \quad \text{for all}\quad t > 0.
\]
\end{lemma}
\comment{
We also prove a curious Siegmund duality relation between $Q^{-\theta}$ and $Q^{2+\theta}$.
\begin{lemma}\label{thm:sdual}
Let $\{X(t), \; t\ge 0\}$ and $\{Y(t), \; t\ge 0\}$ denote a process with law $Q^{-\theta}$ and $Q^{2 + \theta}$ respectively, for some $\theta\ge 0$. Then, for any $x \ge y \ge 0$ we have
\begin{equation}\label{sdual}
Q^{-\theta}_x\left( X(t) \le y \right)= Q^{2+\theta}_y\left( Y(t) \ge x \right).
\end{equation}
\end{lemma}
}
\section{Changing and reversing time}\label{sec:timechange}
Our objective in this section is to establish a time-reversal relationship between NWF and WF models.
\begin{thm}\label{bestimechange}
Let $z_1, \ldots, z_n$ and $\theta_1,\ldots, \theta_n$ be nonnegative constants. Let $Z=(Z_1, \ldots,Z_n)$ be a vector of $n$ independent BESQ processes of dimensions $-\theta_1, \ldots, -\theta_n$ respectively, starting from $(z_1, \ldots, z_n)$. Let $\zeta$ be the sum $\sum_{i=1}^n Z_i$.
Define
\[
T_i= \inf\left\{ t\ge 0: \; Z_i(t)=0 \right\}, \quad \tau=\wedge_{i=1}^n T_i.
\]
Then, there is an $n$-dimensional diffusion $\mu$, satisfying the SDE in \eqref{whatisnwf} for $NWF(\theta_1/2, \ldots, \theta_n/2)$, for which the following equality holds:
\begin{equation}\label{bestime}
Z_i(t\wedge \tau) = \zeta(t\wedge \tau) \mu_i\left( 4C_t \right), \quad 1\le i \le n,\qquad C_t=\int_0^{t\wedge \tau} \frac{ds}{\zeta(s)}.
\end{equation}
Thus, in particular, equation \eqref{whatisnwf} admits a weak solution for all non-negative parameters $(\delta_1, \ldots, \delta_n)$.
\comment{
Moreover, consider the martingale $\beta^*$ defined by
\[
d \zeta(t\wedge \tau) = -\left( \sum_{i=1}^n \theta_i \right) d\left(t \wedge \tau \right) + 2 \sqrt{\zeta(t)} d\beta^*(t).
\]
Then the process $\nu$ is independent of the Dambins-Dubins-Schwarz (DDS) Brownian motion of $\beta^*$ defined over a possibly enlarged space.
}
\end{thm}
\begin{proof} The proof is almost identical to the case of WF model as shown in \cite[Proposition 11]{vsm} with obvious modifications. For example, unlike the WF case, the time-change clock is no longer independent of the NWF process. We outline the basic steps below.
We know from \eqref{sdenbesq} that
\[
d Z_i(t \wedge \tau) = - \theta_i d(t\wedge \tau ) + 2 \sqrt{Z_i} d \beta_i( t \wedge \tau), \quad i=1,2,\ldots,n.
\]
Define $\theta_0=\sum_{i=1}^n \theta_i$. Let $V_i(t)=Z_i/\zeta(t)$ for $t \le \tau$. Then by It\^o's rule we get:
\begin{equation}\label{msde}
dV_i(t\wedge \tau)=-\zeta^{-1}\left[\theta_i - \theta_0 V_i\right]d(t\wedge \tau) + \sqrt{{V_i}(1-V_i)}dM_i(t),
\end{equation}
where
\begin{equation}\label{howtodefmi}
dM_i(t)= \frac{2\zeta^{-1/2}}{\sqrt{1- V_i}}\sum_{j=1}^n \left(1\{i=j\} - \sqrt{V_iV_j}\right)d\beta_j(t\wedge \tau),
\end{equation}
and $\iprod{M_i}(t)=4C_t$.
Let $\{\rho_u, \; u\ge 0 \}$ be the inverse of the increasing function $4C_t$. Applying this time-change to the SDE for $V_i$ in \eqref{msde}, we get
\begin{equation}\label{nuwt}
d\mu_i(t) =- \frac{1}{4}\left[\theta_i - \theta_0 \mu_i\right]dt + \sqrt{\mu_i(1-\mu_i)}\widetilde{W}_i(t),
\end{equation}
where $\widetilde{W}_i$ is the Dambis-Dubins-Schwarz (DDS; see \cite[p.~181]{RY}) Brownian motion associated with $M_i$. This turns out to be the SDE for $NWF(\theta_1/2, \ldots, \theta_n/2)$.
\end{proof}
\comment{
once we prove that the diffusion matrix is given by $\tilde{\sigma}$. To compute it, note that
\[
\begin{split}
\iprod{\mu_i, \mu_j}(4C_t)&= \iprod{V_i, V_j}(t \wedge \tau)= \frac{4}{\zeta(t\wedge \tau)} \sqrt{V_i V_j} \left( 1\{i=j\} - \sqrt{V_iV_j} \right)(t\wedge \tau).
\end{split}
\]
Now changing time by $\rho$, we immediately get $\iprod{\mu_i, \mu_j}=\tilde{\sigma}_{i,j}(\mu)$ as desired.}
Let $\theta_1, \theta_2, \ldots, \theta_n$ be nonnegative and $z_1, z_2, \ldots, z_n$ be positive constants. For $i=1,2,\ldots,n$ define independent random variables $(G_1, \ldots, G_n)$ where $G_i$ is distributed as Gamma($\theta_i/2 + 1$). Let
\begin{equation}\label{whatisri}
R_i=\frac{z_i}{2 G_i}, \qquad i=1,2,\ldots,n.
\end{equation}
Also, independent of $(G_1, \ldots, G_n)$, let $Y_1, Y_2, \ldots, Y_n$ be $n$ independent BESQ processes of positive dimensions $(4+\theta_1),(4+\theta_2), \ldots, (4 + \theta_n)$ respectively, all of which are starting from zero.
For any permutation $\pi$ of $n$ labels, condition on the event
\begin{equation}\label{condperm}
R_{\pi_1} > R_{\pi_2} > \ldots > R_{\pi_n}, \quad \text{and let}\quad R^*=R_{\pi_2}.
\end{equation}
We now construct the following $n$ dimensional process $(X_1, \ldots, X_n)$:
\begin{equation}\label{whatisxi}
X_i(t)=Y_i\left( (t- R^* + R_i)^+ \right), \qquad t\ge 0.
\end{equation}
Notice that at time $t=0$, every $X_i$ is at zero except the $\pi_1$th.
Let $S(t)$ denote the total sum process $\sum_{i=1}^n X_i(t)$. Note that $S(t) >0$ for all $t \ge0$ with probability one. Define the process
\begin{equation}\label{whatisct}
C_t:=\int_{0}^{t}\frac{du}{S(u)},\quad t > 0.
\end{equation}
The process $C_t$ is finite almost surely for every $t$ (unfortunately, we cannot define $R^*=R_{\pi_1}$ precisely because $C_t$ will be infinity; see Lemma \ref{logct}). Let $A$ denote the inverse function of the continuous increasing function $4C$. That is
\begin{equation}\label{whatisa}
A_t = \inf \left\{ u \ge 0:\quad 4C_u \ge t\right\}, \quad t\ge 0.
\end{equation}
\begin{lemma}\label{dimWF} There is an $n$-dimensional diffusion $\nu$ such that the following time-change relationship holds
\begin{equation}\label{getxi}
\nu_i(t)=\frac{X_i}{S}(A_t), \quad \text{or}\quad X_i(t)= S(t) \nu_i\left( 4 C_t \right)\quad t \ge 0.
\end{equation}
The distribution of $\nu$ is supported on the unit simplex
\[
\mathbb{S}_n=\left\{ x_i \ge 0:\quad x_1 + x_2 + \ldots + x_n=1 \right\}.
\]
Conditional on the values of $G_1, \ldots, G_n$ and the process $S$, the law of $\nu$ can be described as below.
Let $\pi$ be any permutation of $n$ labels. On the event $R_{\pi_1} > R^*=R_{\pi_2} > \ldots > R_{\pi_n}$. Let $V_2 < \ldots < V_{n}$ be defined by
\[
A_{V_i}= R^* - R_{\pi_{i}}, \quad \text{or, equivalently}\quad 4C_{R^*-R_{\pi_i}}=V_i.
\]
Note that $V_2=0$.
For $i\ge 2$ and $V_{i} \le t \le V_{i+1}$, the process $\nu$ is zero on all coordinates except $(\pi_1, \ldots, \pi_i)$. The process $\nu(\pi_1, \ldots, \pi_i)$, given the history of the process till time $V_i$ (and the $G_i$'s and $S$) is distributed as the classical Wright-Fisher diffusion starting from
\[
\frac{1}{S}(X_{\pi_1}, \ldots, X_{\pi_i})\left(A_{V_i} \right)=\frac{1}{S}(X_{\pi_1}, \ldots, X_{\pi_i})\left(R^*-R_{\pi_i} \right),
\]
and with parameters $(\gamma_{\pi_1}, \ldots, \gamma_{\pi_i} )$ where
\[
\gamma_j = \theta_j/2 + 2, \quad j=1,2,\ldots,n.
\]
\end{lemma}
\begin{proof} The Gamma random variables $G_1, \ldots, G_n$ are independent of the BESQ process $Y_1, \ldots, Y_n$.
Thus, conditional on $G_1, \ldots, G_n$, the vector of processes $(X_1, \ldots, X_n)$ has the following description. For
\[
R^*- R_{\pi_i}\le t \le R^* - R_{\pi_{i+1}},\quad i\ge 2,
\]
all coordinates other than the $\pi_1$th, $\pi_2$th, $\ldots$, $\pi_i$th are zero. And, $(X_{\pi_1}, \ldots, X_{\pi_i})$, conditioned on the past, are independent BESQ processes of dimensions $(4+\theta_{\pi_1}, \ldots, 4+\theta_{\pi_i})$ and starting from $(X_{\pi_1}, \ldots, X_{\pi_i})(R^*- R_{\pi_i})$.
Thus, on this interval of time, the existence of the process $\nu$, identifying its law as the WF law, and the claimed independence from the process $S$, all follow from \cite[Prop.~11]{vsm}. The proof of the Lemma now follows by combining the argument over the distinct intervals.
\end{proof}
\begin{lemma}\label{timerevdimWF}
Consider the set-up in \eqref{whatisri}, \eqref{whatisxi}, and \eqref{whatisa}. Let $Z_1, Z_2, \ldots, Z_n$ be $n$ stochastic processes defined such that $\{Z_i(t),\; 0\le t \le R^*\}$ is the time-reversal of the process $\{X_i(t),\; 0\le t \le R^*\}$, conditioned on $X_i(R^*)= z_i$. That is, conditioned on $X_i(R^*)=z_i$ for every $i$,
\[
Z_i(t) = X_i(R^*-t)=Y_i(R_i-t)^+, \quad \text{for}\quad 0 \le t \le R^*.
\]
Then $(Z_1, \ldots, Z_n)$ are independent BESQ processes of dimensions $-\theta_1, \ldots, -\theta_n$, starting from $z_1, \ldots, z_n$, and absorbed at the origin.
\end{lemma}
\begin{proof}
It suffices to prove the following:
\medskip
\noindent{\textbf{Claim.}} Let $\{Y(t),\; t\ge 0\}$ denote a BESQ process of dimension $(4+\theta)$ starting from $0$. Fix a $z >0$. Let $T$ be distributed as $z/2G$, where $G$ is a Gamma random variable with parameter $(\theta/2+1)$. Then, conditioned on $T=l$ and $Y(l)=z$, the time-reversed process $\{Y((l-s)^+), 0\le s < \infty\}$ is distributed as $Q_z^{-\theta}$, absorbed at the origin, conditioned on $T_0=l$. Here $T_0$ is the hitting time of the origin for $Q_z^{-\theta}$.
\medskip
Once we prove this claim, the lemma follows since the law of $T_0$ is exactly $z/2G$. See Lemma \ref{lem:trev}.
\medskip
\noindent{\textbf{Proof of Claim.}} For the case of $\theta=0$, this is proved in \cite[p.~447]{besselbridge}. The general proof is exactly similar and we outline just the steps and give references within \cite{besselbridge} for the details.
For any $\theta \in \mathbb{R}$, $t >0$, $x,y \ge 0$, let $Q^{\theta,t}_{x\rightarrow y}$ denote the law of the BESQ bridge of dimension $\theta$, length $t$, from points $x$ to $y$. That is to say, if $Y$ follows $Q_x^\theta$, then $Q^{\theta,t}_{x\rightarrow y}$ is the law of the process $\{ Y(s),\; 0\le s\le t \}$ conditioned on the event $\{ Y(t)=y\}$.
Now, BESQ bridges satisfies time-reversal \cite[p.~446]{besselbridge}. Thus, if we define $\widehat{P}$ to be the $P$-distribution of a process $\{ X(t-s), 0\le s\le t\}$, then $Q^{\theta,t}_{x\rightarrow y}=\widehat Q^{\theta,t}_{y\rightarrow x}$.
We consider the case when the dimension is $(4+\theta), \; \theta \ge 0$, $x=0, y=z >0$. Then
\[
Q^{4+\theta,t}_{0\rightarrow z}=\widehat Q^{4+\theta,t}_{z\rightarrow 0}
\]
Now, from Lemma \ref{scale} (also see \cite[Section 3, p.~440]{besselbridge}) we know that $Q^{4+\theta}_z$ is $Q^{-\theta}_z$ conditioned never to hit zero (or equivalently, $Q^{-\theta}_z$ can be interpreted as $Q_z^{4+\theta}$ conditioned to hit zero). Since the origin is an exit distribution for $Q^{-\theta}_z$ and not an entry (Lemma \ref{scale}; see \cite[p.~441]{besselbridge} for the details of these definitions), the conditional law $Q^{4+\theta,t}_{z\rightarrow 0}$ is nothing but $Q^{-\theta}_z$, conditioned on $T_0=t$. This completes the proof.
\end{proof}
The following is a more precise statement.
Let $(z_1, \ldots, z_n)$ be a point in the $n$-dimensional unit simplex $\mathbb{S}_n$. Fix $n$ nonnegative parameters $\delta_1, \ldots, \delta_n$. Let $G_1, \ldots, G_n$ denote $n$ independent Gamma random variables with parameters $\delta_1+1, \ldots, \delta_n+1$ respectively. Define $R_i=z_i/2G_i$.
For any permutation $\pi$ of $n$ labels, condition on the event $R_{\pi_1} > R_{\pi_2} > \ldots > R_{\pi_n}$, and let $R^*=R_{\pi_2}$.
Define the continuous process $S$ by prescribing $S(0)=Z_1(R_{\pi_1}-R^*)$ where $Z_1$ is distributed as $Q_0^{4+2\delta_{\pi_1}}$; and for any $t$ such that
\[
R^*-R_{\pi_i} \le t \le R^*-R_{\pi_{i+1}},\qquad i\ge 2,\quad R_{\pi_{n+1}}=0.
\]
given the history, the process is distributed as a Bessel-square process of dimension $\sum_{j=1}^i (4+ 2\delta_{\pi_j})$ starting from $S(R^*-R_{\pi_i})$.
Define the stochastic clocks
\[
C_t = \int_{0}^t \frac{du}{S(u)}, \qquad \widehat C_t = \int_{R^*-t}^{R^*} \frac{du}{S(u)},\quad 0\le t\le R^*.
\]
and let $\widehat A_t$ denote the inverse function of $4\widehat C_t$. Let $V_2 < \ldots < V_{n}$ be defined by $4C_{R^*-R_{\pi_i}}=V_i$. Note that $V_2=0$. The $4$ is a standardization constant that appears due to the factor of $2$ in the diffusion coefficient in \eqref{besqintro}.
Define an $n$-dimensional process $\nu$ given $R_1, \ldots, R_n$ and the process $S$.
For $i\ge 2$ and $V_{i} \le t \le V_{i+1}$, the process $\nu$ is zero on all coordinates except possibly at indices $(\pi_1, \ldots, \pi_i)$. At time zero, the process starts at the vector that is $1$ in the $\pi_1$th coordinate and zero elsewhere.
Conditioned on the history till time $V_i$, the process $\{\nu(\pi_1, \ldots, \pi_i)(t),\; V_i \le t \le V_{i+1}\}$ is distributed as the classical Wright-Fisher diffusion starting from $\nu(\pi_1, \ldots, \pi_i)(V_i)$ and with parameters $(\gamma_{\pi_1}, \ldots, \gamma_{\pi_i} )$ where
\[
\gamma_j = \delta_j + 2, \quad j=1,2,\ldots,n.
\]
Finally, consider the conditional law of the process, conditioned on the event
\[
S(R^*) \nu_i(4C_{R^*})= z_i, \quad\text{for all}\quad i=1,2,\ldots,n.
\]
\begin{thm}\label{mainthm}
Define the time-reversed process
\[
\mu(t) = \nu\left( \widehat A \circ 4C_{R^*-t} \right),
\]
where $\circ$ denotes composition. Then this conditional stochastic time-reversed process, until the first time any of the coordinates hit zero, has a marginal distribution (when $G_i$'s and $S$ are integrated out) NWF($\delta_1, \ldots, \delta_n$) starting from $(z_1, \ldots, z_n)$.
\end{thm}
\comment{
This leads to the proof of Theorem \ref{mainthm}.
\begin{thm} Let $z_1, \ldots, z_n$ and $R_1, \ldots, R_n$ be as in \eqref{whatisri}. Let us condition on the event $R_{\pi_1} > R_{\pi_2} > \ldots > R_{\pi_n}$. Define the continuous process $S(t)$ given by $S(0)=0$; and for $R^*-R_{\pi_i} \le t \le R^*-R_{\pi_{i+1}}$, given the history, the process is distributed as BESQ of dimension $\sum_{j=1}^i (4+\theta_j)$ starting from $S(R^*-R_{\pi_i})$.
Define the stochastic clocks
\[
C_t = \int_0^t \frac{du}{S(u)}, \qquad \widehat C_t = \int_{R^*-t}^{R^*} \frac{du}{S(u)},\quad 0\le t\le R^*.
\]
and let $\widehat A_t$ denote the inverse function of $\widehat C_t$.
Consider the process $\nu$ as in Lemma \ref{dimWF} and consider the conditional law of the process, conditioned on the event \[
\nu_i(C_{R^*})= \frac{z_i}{z_1+\ldots + z_n}, \quad\text{for all}\quad i=1,2,\ldots,n.
\]
Define the time-reversed process
\[
\mu(t) = \nu\left( \widehat A \circ C_{R^*-t} \right),
\]
where $\circ$ denotes composition.
Then this conditional stochastic time-changed process is distributed as NWF with parameters ($-\theta_1/2, -\theta_2/2, \ldots, -\theta_n/n$), starting from $(z_1, \ldots, z_n)$ until the first time any of the coordinates hit zero.
\end{thm}
}
\begin{proof}[Proof of Theorem \ref{mainthm}]
We start with given values of $R_{\pi_1} > R_{\pi_2} > \ldots > R_{\pi_n}$ and the process $S$. Applying equation \eqref{getxi} in Lemma \ref{dimWF} to obtain the processes $(X_1, \ldots, X_n)$ defined by
\[
X_i(t) = S(t) \nu_i(4C_t), \quad 0\le t \le R^*.
\]
Then, the vector $(X_1, X_2, \ldots, X_n)$ has the law prescribed by \eqref{whatisxi}.
Now we apply Lemma \ref{timerevdimWF} to obtain $(Z_1, \ldots, Z_n)$ by conditioning $(X_1, \ldots, X_n)$ and reversing time.
Finally the construction in Theorem \ref{bestimechange} gives us the vector $(\mu_1, \ldots, \mu_n)$ from $(Z_1, \ldots, Z_n)$ as desired.
\end{proof}
\section{Exit density}\label{sec:exitdensity}
Let $Z_1, Z_2, \ldots, Z_n$ be independent BESQ processes of dimensions $-\theta_1, \ldots, -\theta_n$, where each $\theta_i \ge 0$. We assume that at time zero, the vector $\mathbf{Z}=(Z_1, \ldots, Z_n)$ starts from a point $\mathbf{z}=(z_1, \ldots, z_n)$ where every $z_i > 0$. Define $T_i$ to be the first hitting time of zero for the process $Z_i$, and let $\tau=\wedge_{i} T_i$ denotes the first time any coordinate hits zero. We would like to determine the joint distribution of $(\tau, \mathbf{Z}(\tau))$.
Note that since each $T_i$ is a continuous random variable, the minimum is attained at a unique $i$. Thus, for a fixed $1\le i \le n$, conditioned on the event $\tau=T_i$, the distribution of $Z_i(\tau)$ is the unit mass at zero, and the distribution of every other $Z_j(\tau)$ is supported on $(0, \infty)$. Now, let $h_i$ denote the density of the stopping time $T_i$ on $(0,\infty)$ and let $q_t^{-\theta}$ refers to the transition density of $Q^{-\theta}$. It follows that for any $a_j > 0$, $j\neq i$, we get
\[
\begin{split}
P&\left( \tau=T_i, \tau\le t, Z_j(\tau) \ge a_j, \;\text{for all $j \neq i$} \right)\\
=&P\left( T_i \le t, T_j > T_i, Z_j(T_i) \ge a_j, \; \text{for all $j\neq i$} \right)\\
=& \int_0^t h_i(s) \prod_{j\neq i} P\left( T_j > s, Z_j(s) \ge a_j \right)ds= \int_0^t h_i(s) \prod_{j\neq i}P\left( Z_j(s) \ge a_j \right)ds, \quad \text{since $a_j > 0$.}\\
=& \int_0^t h_i(s) \left[ \prod_{j\neq i} \int_{a_j}^\infty q_s^{-\theta_j}(z_j, y_j) dy_j \right] ds.
\end{split}
\]
Our first job is to find closed form expressions of the integral above. To do this we start by noting that $T_i$ is distributed as $z_i/ 2G_i$ (see Lemma \ref{lem:trev}), where $G_i$ is a Gamma random variable with parameter $(4+\theta_i)/2-1=\theta_i/2 + 1$. That is, the density of $G_i$ is supported on $(0,\infty)$ and is given by
\[
\frac{y^{\theta_i/2}}{\Gamma(\theta_i/2 + 1)}e^{-y}.
\]
It follows that
\[
h_i(s)= \frac{(z_i/2)^{\theta_i/2+1}}{\Gamma(\theta_i/2 + 1)} s^{-\theta_i/2 -2} e^{-z_i/2s}, \quad 0 \le s < \infty.
\]
On the other hand, it follows from time reversal (Lemma \ref{lem:trev}) that $q_s^{-\theta_j}(z_j, y_j)= q_{s}^{4+\theta_j}(y_j, z_j)$. For any positive $a$, the transition density $q^{a}_s(y,z)$ is explicitly known (see, e.g., in \cite{vsm}) to be $s^{-1}f(z/s, a, y/s)$, where $f(\cdot,k,\lambda)$ is the density of a noncentral Chi-square distribution with $k$-degrees of freedom and a noncentrality parameter value $\lambda$. In particular, it can be written as a Poisson mixture of central Chi-square (or, Gamma) densities. Thus, we have the following expansion
\begin{equation}\label{tranden}
q_s^{-\theta_j}(z_j, y_j)= q_{s}^{4+\theta_j}(y_j, z_j)=s^{-1}\sum_{k=0}^{\infty} e^{-y_j/2s}\frac{(y_j/2s)^k}{k!} g_{\theta_j+4+2k}(z_j/s),
\end{equation}
where $g_r$ is the Gamma density with parameters $(r/2, 1/2)$. That is
\[
g_r(x)= \frac{2^{-r/2} x^{r/2-1}}{\Gamma(r/2)} e^{-x/2}, \quad x\ge 0.
\]
Now, define
\[
\vec{y}_i=\sum_{j\neq i} y_j, \quad \vec{\theta}_i= \sum_{j\neq i}\theta_j,\quad \vec{z}_i=\sum_{j\neq i} z_j
\]
Thus
\[
\begin{split}
&h_i(s)\prod_{j\neq i} q_s^{-\theta_j}(z_j, y_j)= \frac{(z_i/2)^{\theta_i/2+1}}{\Gamma(\theta_i/2 + 1)} s^{-\theta_i/2 -2} e^{-z_i/2s} \prod_{j\neq i} s^{-1}\sum_{k=0}^{\infty} e^{-y_j/2s}\frac{(y_j/2s)^k}{k!} g_{\theta_j+4+2k}(z_j/s)\\
&=\frac{(z_i/2)^{\theta_i/2+1}}{\Gamma(\theta_i/2 + 1)} s^{-\theta_i/2 -2} e^{-z_i/2s} \prod_{j\neq i} s^{-1}\sum_{k=0}^{\infty} e^{-y_j/2s}\frac{(y_j/2s)^k}{k!} \frac{2^{-\theta_j/2-2-k} (z_j/s)^{\theta_j/2+k+1}}{\Gamma(\theta_j/2 + 2 +k)} e^{-z_j/2s}\\
&=\frac{(z_i/2)^{\theta_i/2+1}}{\Gamma(\theta_i/2 + 1)} s^{-\theta_i/2 - 2-(n-1)} e^{-z_i/2s}\times \\
& e^{-(\vec{y}_i+ \vec{z}_i)/2s}2^{-\vec{\theta}_i/2-2(n-1)} \prod_{j\neq i} \sum_{k=0}^{\infty} \frac{(y_j/2s)^k}{k!} \frac{2^{-k} (z_j/s)^{\theta_j/2+k+1}}{\Gamma(\theta_j/2 + 2 +k)}.
\end{split}
\]
We now exchange the product and the sum in the above. We will need some more notations for a compact representation.
For any two vectors $a$ and $b$, denote by
\[
a^b = \prod_{i} a_i^{b_i}, \quad a!=\prod_{i} a_i!.
\]
Also let $\mathbf{\Theta}_i, \mathbf{y}_i, \mathbf{z}_i$ stand for the vectors $(\theta_j,\; j\neq i)$, $(y_j, \; j\neq i)$, and $(z_j, \; j\neq i)$ respectively.
Let $\mathbf{k}$ denote the vector $(k_j,\; j\neq i)$, where every $k_j$ takes any nonnegative integer values. Let $\mathbf{k}'1$ be the sum of the coordinates of $\mathbf{k}$. Then
\[
\begin{split}
\prod_{j\neq i} &\sum_{k=0}^{\infty} \frac{(y_j/2s)^k}{k!} \frac{2^{-k} (z_j/s)^{\theta_j/2+k+1}}{\Gamma(\theta_j/2 + 2 +k)} \\
&= \sum_{N=0}^{\infty} (4s)^{-N} s^{-\vec{\theta}_i/2 - N - (n-1)} \mathbf{z}_i^{\mathbf{\Theta}_i/2+1}\sum_{\mathbf{k}'1=N} \frac{\mathbf{y}_i^\mathbf{k}}{\mathbf{k}!} \frac{\mathbf{z}_i^{\mathbf{k}}}{\prod_{j\neq i} \Gamma(\theta_j/2 + 2 + k_j)}.
\end{split}
\]
Thus, combining the expressions we get
\begin{equation}\label{inter1}
\begin{split}
h_i(s)&\prod_{j\neq i} q_s^{-\theta_j}(z_j, y_j)=\frac{z_i^{\theta_i/2+1}}{\Gamma(\theta_i/2 + 1)}2^{-\theta_i/2 -1 -\vec{\theta}_i/2-2(n-1)}\times \\
& s^{-\theta_i/2 - 2-(n-1)} e^{-z_i/2s} e^{-(\vec{y}_i+ \vec{z}_i)/2s} \sum_{N=0}^{\infty} 4^{-N} s^{-\vec{\theta}_i/2 - 2N - (n-1)}B_N,
\end{split}
\end{equation}
where
\[
B_N= \mathbf{z}_i^{\mathbf{\Theta}_i/2+1}\sum_{\mathbf{k}'1=N} \frac{\mathbf{y}_i^\mathbf{k}}{\mathbf{k}!} \frac{\mathbf{z}_i^{\mathbf{k}}}{\prod_{j\neq i} \Gamma(\theta_j/2 + 2 + k_j)}.
\]
We can now integrate over $s$ in \eqref{inter1} to obtain
\[
\begin{split}
\int_0^\infty h_i(s)&\prod_{j\neq i} q_s^{-\theta_j}(z_j, y_j)ds= \sum_{N=0}^\infty B'_N \int_0^\infty s^{-a_N} e^{-b/s} ds,
\end{split}
\]
where
\begin{eqnarray}
B_N' &=& \frac{z_i^{\theta_i/2+1}}{\Gamma(\theta_i/2 + 1)}2^{-\theta_0/2-2n+1} 4^{-N} B_N, \quad \theta_0=\sum_{i=1}^n \theta_i,\\
a_N &=& \theta_i/2 + \vec{\theta}_i/2 +2n + 2N= \theta_0/2 + 2n + 2N,\\
b &=& z_i/2 + (\vec{y}_i+\vec{z}_i)/2 = (\vec{y}_i + z_0)/2, \quad z_0=\sum_{i=1}^n z_i.
\end{eqnarray}
Now a simple change of variable $w=1/s$ shows
\[
\begin{split}
\int_0^\infty s^{-a_N} e^{-b/s}ds &= \int_0^\infty w^{a_N}e^{-bw}w^{-2}dw=\int_0^\infty w^{a_N -2} e^{-bw}dw\\
\frac{\Gamma(a_N-1)}{b^{a_N-1}}&\int_0^{\infty} \frac{b^{a_N-1}}{\Gamma(a_N-1)} w^{a_N -2} e^{-bw}dw= \frac{\Gamma(a_N-1)}{b^{a_N-1}}.
\end{split}
\]
Since the $i$th coordinate of the exit point is zero, one can define $y_i=0$ and $y_0=\sum_{j=1}^n y_j= \vec{y}_i$ to simplify notation.
Thus we obtain
\[
\begin{split}
\int_0^\infty &h_i(s)\prod_{j\neq i} q_s^{-\theta_j}(z_j, y_j)ds = \sum_{N=0}^\infty \frac{z_i^{\theta_i/2+1}}{\Gamma(\theta_i/2 + 1)}2^{-\theta_0/2-2n+1} 4^{-N}B_N \frac{\Gamma(a_N-1)}{b^{a_N-1}}\\
&=\frac{z_i^{\theta_i/2+1}\mathbf{z}_i^{\mathbf{\Theta}_i/2+1}}{\Gamma(\theta_i/2 + 1)}2^{-\theta_0/2-2n+1} \sum_{N=0}^\infty \left( (\vec{y}_i+z_0)/2 \right)^{-\theta_0/2 - 2n - 2N + 1}\\
&\times \Gamma(\theta_0/2 + 2n + 2N-1) 4^{-N}\sum_{\mathbf{k}'1=N} \frac{\mathbf{y}_i^\mathbf{k}}{\mathbf{k}!} \frac{\mathbf{z}_i^{\mathbf{k}}}{\prod_{j\neq i} \Gamma(\theta_j/2 + 2 + k_j)}\\
&=\frac{\mathbf{z}^{\mathbf{\Theta}/2+1}}{\Gamma(\theta_i/2 + 1)}2^{-\theta_0/2-2n+1} \sum_{N=0}^\infty \left( y_0 + z_0 \right)^{-\theta_0/2 - 2n - 2N + 1} 2^{\theta_0/2 + 2n + 2N -1}\\
&\times \Gamma(\theta_0/2 + 2n + 2N-1) 4^{-N}\sum_{\mathbf{k}'1=N} \frac{\mathbf{y}_i^\mathbf{k}}{\mathbf{k}!} \frac{\mathbf{z}_i^{\mathbf{k}}}{\prod_{j\neq i} \Gamma(\theta_j/2 + 2 + k_j)}\\
&=\frac{\mathbf{z}^{\mathbf{\Theta}/2+1}}{\Gamma(\theta_i/2 + 1)} \sum_{N=0}^\infty \left( y_0 + z_0 \right)^{-\theta_0/2 - 2n - 2N + 1}\times \\
&\Gamma(\theta_0/2 + 2n + 2N-1)\sum_{\mathbf{k}'1=N} \frac{\mathbf{y}_i^\mathbf{k}}{\mathbf{k}!} \frac{\mathbf{z}_i^{\mathbf{k}}}{\prod_{j\neq i} \Gamma(\theta_j/2 + 2 + k_j)}.
\end{split}
\]
We have the following result.
\begin{thm}\label{thm:besexit}
Let $Z_1, Z_2, \ldots, Z_n$ be independent BESQ processes of dimensions $-\theta_1, \ldots, -\theta_n$, where each $\theta_i\ge 0$. Assume that $Z_i(0)= z_i(0) > 0$, for every $i$.
The distribution of $(\tau, Z({\tau}))$ is supported on the set $(0,\infty)\times \cup_{i=1}^n H_i$, where $H_i$ is the subspace orthogonal to the $i$th canonical basis vector $e_i$. That is
\[
H_i = \left\{ (y_1, y_2, \ldots, y_n): \; y_i=0 \right\}.
\]
\begin{enumerate}
\item[(i)] Let $G_i,\; i=1,2,\ldots n$ be independent Gamma random variables with parameters $\theta_i/2 + 1, \; i=1,2,\ldots, n$.
The law of $\tau$ is the same as that of $\min_{i} \frac{z_i}{2G_i}$ and
\[
P\left( \tau=T_i \right)= P\left( \frac{G_i}{z_i} > \frac{G_j}{z_j}, \; \text{for all $j\neq i$} \right),
\]
where $T_i$ is the first hitting time of $H_i$.
\item[(ii)] The restriction of the law of the random vector $Z(\tau)$, restricted to the hyperplane $H_i$, admits a density with respect to all the variables $y_j$'s, $j\neq i$, which is given by
\begin{equation}\label{eq:eden}
\begin{split}
&=\frac{S^{1-\theta_0/2-2n}}{\Gamma(\theta_i/2 + 1)}\prod_{j=1}^n z_j^{\theta_j/2 + 1}\sum_{N=0}^\infty \Gamma(\theta_0/2 + 2n + 2N-1) S^{- 2N}\\
&\times\sum_{\sum_{j\neq i} k_j=N} \prod_{j\neq i} \frac{(y_j z_j)^{k_j}}{k_j! \Gamma(\theta_j/2 + 2 + k_j)}.
\end{split}
\end{equation}
Here
\[
S=\sum_{i=1}^n (y_i + z_i), \quad y_i=0, \quad \theta_0=\sum_{i=1}^n \theta_i.
\]
\end{enumerate}
\end{thm}
Using Theorem \ref{bestimechange}, we get that the exit distribution of $NWF(\delta_1, \ldots, \delta_n)$ starting from a point $(z_1, \ldots, z_n)\in \mathbb{S}_n$ is the image under the map
\[
x_i \mapsto \frac{x_i}{\sum_{j=1}^n x_j}, \qquad 1\le i\le n
\]
of the exit density of independent BESQ processes of dimensions $-\theta_1, \ldots, -\theta_n$, where each $\theta_i=2\delta_i$.
\begin{thm}\label{thm:exitnwf}
The exit density of $\mu \sim$ NWF$(\delta_1, \ldots, \delta_n)$ starting from $(z_1, \ldots, z_n)\in \mathbb{S}_n$ is supported on the set $\cup_{i=1}^n F_i$, where $F_i$ is the face $\{ x \in \mathbb{S}_n:\; x_i=0 \}$, and admits the following description.
\begin{enumerate}
\item[(i)] Let $G_i,\; i=1,2,\ldots n$ be independent Gamma random variables with parameters $\delta_i + 1, \; i=1,2,\ldots, n$.
Then
\begin{equation}\label{eq:exithyp}
P\left( \text{$\mu$ exits through $F_i$} \right)= P\left( \frac{G_i}{z_i} > \frac{G_j}{z_j}, \; \text{for all $j\neq i$} \right),
\end{equation}
\item[(ii)] Let $\delta$ represent the vector $(\delta_1, \ldots, \delta_n)$ and let $\delta_0=\sum_{i=1}^n \delta_i$. The exit distribution of the process $\mu$, restricted to $F_i$, admits a density with respect to all the variables $x_j$'s, $j\neq i$, which is given by
\begin{equation}\label{eq:exitnwf}
\begin{split}
(\delta_i + 1) \sum_{N=0}^\infty \frac{\Gamma(N+n+\delta_0)}{\Gamma(N+2n + \delta_0)} \sum_{\sum_{j\neq i} k_j=N} \text{Dir}_n(z;\mathbf{k} + \delta +\mathbf{2}) \text{Dir}_{n-1}(x;\mathbf{k}+\mathbf{1}).
\end{split}
\end{equation}
Here the inner sum above is over all nonnegative integers $(k_j, \; j \neq i)$ such that $\sum_{j\neq i}k_j =N$. The vector $\mathbf{k}$ represents a vector whose $j$th coordinate is $k_j$ for all $j \neq i$, and $\mathbf{k}_i=0$. The vectors $\mathbf{k}+\delta+\mathbf{2}$ and $\mathbf{k}+\mathbf{1}$ represent vector additions of $\mathbf{k}$, $\delta$, and the vector of all twos, and $\mathbf{k}$ and the vector of all ones respectively. The factor $\text{Dir}_{n-1}$ is a density with respect to the $(n-1)$-dimensional vector $(x_j, \; j\neq i)$ with corresponding parameters $(\mathbf{k}_{j}+1,\; j \neq i)$. It can also be interpreted as the conditional density of the $n$-dimensional $\text{Dir}_n(x;\mathbf{k}+1)$, conditioned on $x_i=0$.
\end{enumerate}
\end{thm}
Note that the density in \eqref{eq:exitnwf} is a mixture of Dirichlet densities, strikingly similar to those appearing as transition probabilities of the Wright-Fisher diffusions themselves. See Griffiths \cite{griffiths79b}, Barbour, Ethier, and Griffiths \cite{BEG}, and Pal \cite{vsm}.
\begin{proof}[Proof of Theorem \ref{thm:exitnwf}] This is a straightforward integration. We have assumed that $\sum_i z_i=1$. Thus, $S=1+\sum_j y_j$; define $y_0=\sum_j y_j$, and
\[
x_j = y_j/y_0, \quad 1\le j \le n.
\]
Hence \eqref{eq:eden} simplifies to
\begin{equation}
\begin{split}
&=\frac{(1+y_0)^{1-\theta_0/2-2n}}{\Gamma(\theta_i/2 + 1)}\prod_{j=1}^n z_j^{\theta_j/2 + 1}\sum_{N=0}^\infty \Gamma(\theta_0/2 + 2n + 2N-1) (1+y_0)^{-2 N} \\
&\times y_0^N\sum_{\sum_{j\neq i} k_j=N} \prod_{j\neq i} \frac{(x_j z_j)^{k_j}}{k_j! \Gamma(\theta_j/2 + 2 + k_j)}.
\end{split}
\end{equation}
Now, to get to formula \eqref{eq:exitnwf} we need to make a multivariate change of variables. Without loss of generality, let $i=n$. Then, for any $y\in F_i$, we have $y_{n}=0$. Define the change of variables
\[
(y_1, \ldots, y_{n-2}, y_{n-1}) \mapsto \left( y_0,x_1, \ldots, x_{n-2}\right).
\]
In other words, $y_i= y_0 x_i$, for all $i=1,2,\ldots, n-2$ and $y_{n-1}=y_0(1-x_1-\ldots - x_{n-2})$. The determinant of the well-known Jacobian matrix is given by $y_0^{n-2}$.
Thus, the density of $(x_1, \ldots, x_n)$ restricted to $F_i$ is given by
\begin{equation}\label{denform1}
\begin{split}
\frac{1}{\Gamma(\theta_i/2 + 1)}&\prod_{j=1}^n z_j^{\theta_j/2 + 1}\sum_{N=0}^\infty \Gamma(\theta_0/2 + 2n + 2N-1)\\
\int_0^\infty &y^{N+n-2}(1+y)^{1-\theta_0/2-2n-2N}dy \times\sum_{\sum_{j\neq i} k_j=N} \prod_{j\neq i} \frac{(x_j z_j)^{k_j}}{k_j! \Gamma(\theta_j/2 + 2 + k_j)}.
\end{split}
\end{equation}
The following formula is easily verifiable for $\alpha \ge 0$, $\beta > \alpha+1$:
\[
\int_0^\infty y^{\alpha} (1+y)^{-\beta} dy=\int_0^1 x^{\beta-\alpha -2}(1-x)^{\alpha} dx=B(\alpha+1,\beta-\alpha-1)
\]
where $B$ refers to the Beta function.
In other words, \eqref{denform1} reduces to
\begin{equation}\label{denform2}
\begin{split}
&\frac{1}{\Gamma(\theta_i/2 + 1)}\prod_{j=1}^n z_j^{\theta_j/2 + 1}\sum_{N=0}^\infty \Gamma(\theta_0/2 + 2n + 2N-1)\\
& B(N+n-1, N+n+\theta_0/2) \sum_{\sum_{j\neq i} k_j=N} \prod_{j\neq i} \frac{(x_j z_j)^{k_j}}{k_j! \Gamma(\theta_j/2 + 2 + k_j)}.
\end{split}
\end{equation}
We now change $\theta_i/2$ to $\delta_i$ and rewrite the above expression in terms of Dirichlet densities. We use the notations in the statement of Theorem \ref{thm:exitnwf}: the vector $\mathbf{k}$ represents a vector whose $j$th coordinate is $k_j$ for all $j \neq i$, and $\mathbf{k}_i=0$. The vectors $\mathbf{k}+\delta+\mathbf{2}$ and $\mathbf{k}+\mathbf{1}$ represent vector additions of $\mathbf{k}$, $\delta$, and the vector of all twos, and $\mathbf{k}$ and the vector of all ones respectively. The factor $\text{Dir}_{n-1}$ is a density with respect to the $(n-1)$-dimensional vector $(x_j, \; j\neq i)$ with corresponding parameters $(\mathbf{k}_{j}+1,\; j \neq i)$. It can also be interpreted as the conditional density of the $n$-dimensional $\text{Dir}_n(x;\mathbf{k}+1)$, conditioned on $x_i=0$.
Hence, for any $(k_j, \; j\neq i)$, integers,
\[
\begin{split}
\frac{z_i^{\delta_i+1}}{\Gamma(\delta_i+1)}&\prod_{j\neq i} \frac{z_j^{k_j+\delta_j+1}}{\Gamma(\delta_j + 2 + k_j)} \frac{x_j^{k_j}}{k_j!}=\\
& \frac{(\delta_i+1)}{\Gamma(\delta_0+N+2n)\Gamma(N+n-1)} \text{Dir}_n(z;\mathbf{k}+\delta+\mathbf{2}) \text{Dir}_{n-1}(x; \mathbf{k}+\mathbf{1}).
\end{split}
\]
Thus \eqref{denform2} reduces to
\begin{equation}\label{denform3}
\begin{split}
(\delta_i+1)& \sum_{N=0}^\infty \frac{\Gamma(\delta_0 + 2n + 2N-1) B(N+n-1, N+n+\delta_0) }{\Gamma(\delta_0+N+2n)\Gamma(N+n-1)}\\
&\times \sum_{\mathbf{k}'\mathbf{1}=N} \text{Dir}_n(z;\mathbf{k}+\delta+\mathbf{2}) \text{Dir}_{n-1}(x; \mathbf{k}+\mathbf{1}).
\end{split}
\end{equation}
However
\[
\begin{split}
&\frac{\Gamma(\delta_0 + 2n + 2N-1) B(N+n-1, N+n+\delta_0) }{\Gamma(\delta_0+N+2n)\Gamma(N+n-1)}\\
&= \frac{\Gamma(\delta_0 + 2n + 2N-1)}{\Gamma(\delta_0+N+2n)\Gamma(N+n-1)}\frac{\Gamma(N+n-1)\Gamma(N+n+\delta_0)}{\Gamma(2N+2n+\delta_0-1)}\\
&=\frac{\Gamma(N+n+\delta_0)}{\Gamma(N+2n+\delta_0)}.
\end{split}
\]
This completes the proof of formula \eqref{eq:exitnwf}.
The probability in \eqref{eq:exithyp} is a direct consequence of Theorem \ref{thm:besexit} conclusion (i).
\end{proof}
\section{Exit time}\label{sec:exittime}
Let $X=(X_1, \ldots, X_n)$ be distributed as NWF$(-\theta_1/2, \ldots, -\theta_n/2)$ starting from a point $(x_1, \ldots, x_n)$ in the unit simplex. Let $\sigma_0$ denote the stopping time
\[
\sigma_0 = \inf \left\{ t\ge 0:\; X_i =0\quad \text{for some $i$} \right\}.
\]
Our objective is to find estimates on the law of $\sigma_0$.
We will simplify the situation by assuming that all $x_i=1/n$ and all $\theta_i=\theta$. To this end we use the time-change relationship in Theorem \ref{bestimechange}. Let $Z=(Z_1, \ldots, Z_n)$ be independent BESQ processes starting from $(z_1, \ldots, z_n)$ as in the set-up of Theorem \ref{bestimechange}, where each $z_i$ is now one. Then
\begin{equation}\label{whatissigma0}
\sigma_0 = 4\int_0^{\tau} \frac{ds}{\zeta(s)}, \qquad \zeta(s)=\sum_{i=1}^n Z_i(s).
\end{equation}
By Theorem \ref{thm:besexit}, the distribution of $\tau$ is the same as considering $n$ iid Gamma$(\theta/2 +1)$ random variables $G_1, \ldots, G_n$, and defining
\begin{equation}\label{whatistau}
\tau=\frac{1}{2\max_i G_i}.
\end{equation}
Our first step will be to prove a concentration estimate of $\max_i G_i$.
\begin{lemma}\label{maxmom}
Let $G_1, G_2, \ldots, G_n$ be $n$ iid Gamma random variables with parameter $r/2$, for some $r\ge 2$. Let $\chi$ be the random variable $\max_i G_i$.
Then, as $n$ tends to infinity,
\[
E\sqrt{\chi} = \Theta\left( \sqrt{\log n} \right).
\]
\end{lemma}
\begin{proof} First let $r\in \mathbb{N}$. Let $\{Z_1(i), \ldots, Z_n(i), \; i=1,2,\ldots,r \}$ be a collection of iid standard Normal random variables. Then $2G_j$ has the same law as $Z_j^2(1) + \ldots + Z_j^2(r)$. Hence
\[
E \max_j \abs{Z}_j(1) \le E \sqrt{2\chi} \le \sqrt{r} E \max_{i,j} \abs{Z}_j(i)
\]
As $n$ tends to infinity, the right hand side above converges to $\sqrt{2r\log (rn)}$ while the left hand side converges to $\sqrt{2\log n}$. This completes the argument for $r\in \mathbb{N}$. For a general positive $r$, bound on both sides by $\lfloor r \rfloor$ and $\lfloor r \rfloor + 1$.
\end{proof}
\comment{
\begin{lemma}\label{maxmom}
Let $G_1, G_2, \ldots, G_n$ be $n$ iid Gamma random variables with parameter $r\ge 1$. Let $\chi$ be the random variable $\max_i G_i$. Then
Let $\varpi(x)$ denote the functional inverse of $\Gamma(r) x^{1-r}e^{x}$. Then
\[
\chi = a_n + b_n O_p(1),
\]
where $a_n \sim \varpi(n)$, $b_n= o(a_n)$, and $O_p(1)$ refers to a tight sequence of random variables.
\end{enumerate}
\end{lemma}
\begin{proof}
Let $G$ be a Gamma random variable with parameter $r \ge 1$. Then
\[
E(e^{\lambda G}) = \frac{1}{\Gamma(r)} \int_0^\infty e^{\lambda y} y^{r-1} e^{-y} dy < \infty, \quad \text{since}\; \lambda < 1.
\]
Now $\max_i G_i < G_1 + \ldots + G_n$. Since each $G_i$ has finite exponential moment for $0 < \lambda < 1$, so does $\max_i G_i$. The case when $\lambda \le 0$ is trivial. This proves (i).
For part (ii) we use some standard results from Extreme Value Theory. We refer the reader to the book by de Haan and Ferreira \cite{dHF} for all the theorems we use below. Let $F$ denote the distribution function of a Gamma($r$) random variable. Then
$F'(x) = \frac{1}{\Gamma(r)} x^{r-1} e^{-x}$ is the Gamma density. The asymptotics of the function $1-F$ (as $x$ tends to infinity) can be obtained from the asymptotics of the incomplete gamma function which are listed in \cite{AS}.
\begin{equation}\label{gammaasy}
\Gamma(r)(1 - F(x)):= \Gamma(r,x) \sim x^{r-1} e^{-x}\left[ 1 + \frac{r-1}{x} + \frac{(r-1)(r-2)}{x^2} + \ldots \right], \quad x \rightarrow \infty.
\end{equation}
We now apply the \textit{Von Mises} condition \cite[p.~15]{dHF} and evaluate
\[
\begin{split}
\lim_{x\rightarrow \infty}\frac{d}{dx}\left( \frac{1-F}{F'} \right)(x) &= \lim_{x\rightarrow \infty} \frac{d}{dx} e^x x^{1-r}\int_x^\infty t^{r-1}e^{-t}dt\\
&=\lim_{x\rightarrow\infty}\left[ \frac{\Gamma(r,x)}{x^{r-1}e^{-x}} - 1 + \frac{1-r}{x} \frac{\Gamma(r,x)}{x^{r-1}e^{-x}}\right]=0.
\end{split}
\]
The final equality above is in view of \eqref{gammaasy}.
Hence by \cite[Theorem 1.1.8]{dHF} we get that there are two sequences $\{ a_n, b_n\}$ such that
\[
\lim_{n\rightarrow \infty }P\left( \frac{\chi - a_n}{b_n} \le x \right) = e^{-e^{-x}}, \quad x > 0.
\]
Moreover, let $U(t)$ to be the functional inverse of $1/(1-F)$, i.e.,
\[
U(t) = \inf\left\{ x\ge 0:\; \frac{1}{1-F(x)} > t \right\}=\left\{ x: F(x) = 1 - \frac{1}{t}\right\}.
\]
Then $a_n=U(n)$, $b_n=nU'(n)$ and (\cite[p.~22]{dHF}) $b_n = o(a_n)$.
From \eqref{gammaasy} we get that $1-F(x)$ is asymptotically of the same order as the functional inverse of $x^{r-1}e^{-x}/\Gamma(r)$. Since all our functions are strictly increasing, the inverse of $1/(1-F(x))$ is asymptotically of the same order as the inverse of $\Gamma(r)x^{1-r}e^{x}$. This completes the proof.
\end{proof}
}
We also need a version of logarithmic Sobolev inequality for Gamma random variables which can be found in several articles including \cite{BW}.
\begin{lemma}{\cite[p.~2718]{BW}}\label{lem:logsobo}
Let $\mu^\theta$ denote the product probability measure of $n$ iid Gamma$(\theta)$ random variables. Then, for every $f$ on $\mathbb{R}^n$ which is in $C^1$ (i.e., once continuously differentiable) one has
\begin{equation}\label{logsobo}
\mathrm{Ent}(f^2) \le 4 \int \left( \sum_{i=1}^n x_i \left( \partial_i f(x) \right)^2 \right)d\mu^{\theta}(x).
\end{equation}
Here $\mathrm{Ent}(\cdot)$ refers to the entropy defined by
\[
\mathrm{Ent}(f^2)=\int f^2 \log(f^2) d\mu^\theta - \left( \int f^2 d\mu^\theta \right) \log \left(\int f^2 d\mu^\theta \right).
\]
And $\partial_i$ refers to the partial derivative with respect to the $i$th coordinate.
\end{lemma}
\begin{lemma}\label{expconcen}
Consider the set-up in Lemma \ref{lem:logsobo}. Let $F$ be a function on the open positive quadrant (i.e., every $x_i>0$) which is $C^1$ and satisfies
\begin{equation}\label{derivbnd}
\sum_{i=1}^n x_i \left( \partial_i F \right)^2 \le F.
\end{equation}
Then the following concentration estimate holds for any $r >0$:
\[
\begin{split}
\mu^\theta\left( \sqrt{F} - E_\theta \sqrt{F} \ge r \right) &\le \exp\left( -r^2 \right),\quad \mu^\theta\left( \sqrt{F} - E_\theta \sqrt{F} \le - r \right) \le \exp\left( -r^2 \right)\\
\end{split}
\]
where $E_{\theta}\sqrt F = \int \sqrt{F} d\mu^\theta$.
\end{lemma}
\begin{proof} Condition \eqref{derivbnd} implies that $4\sum_{i=1}^n x_i \left( \partial_i \sqrt{F}\right)^2 \le 1$.
Hence, from the classical Herbst argument (for example, the monograph by Ledoux \cite{L}), with a gradient defined by the right side of \eqref{logsobo} we get
\[
\mu^\theta\left( \sqrt{F} -E_\theta \sqrt{F} > r \right) \le \exp\left(- r^2 \right).
\]
Here $\mu^\theta(\sqrt{F})$ is the expectation of $\sqrt{F}$ under $\mu^\theta$. Repeating the argument with $-\sqrt{F}$ instead of $\sqrt{F}$ we get the result.
\end{proof}
\comment{
We follow a line of argument similar to the classical Herbst argument. The classical argument can be found in, for example, the monograph by Ledoux \cite{L}.
Let $F$ be as in the statement of the theorem and apply Lemma \ref{lem:logsobo} to
\[
f= e^{\lambda F/2}, \quad \lambda < 1.
\]
Note that
\[
\sum_{i=1}^n x_i \left( \partial_i f(x) \right)^2 =\frac{\lambda^2}{4} \sum_{i=1}^n x_i \left( \partial_i F \right)^2 e^{\lambda F} \le \frac{\lambda^2}{4} F e^{\lambda F}.
\]
Define $H(\lambda) = E_{\theta}(e^{\lambda F})$, $\lambda \in (0,1)$. Apply inequality \eqref{logsobo} we get
\[
\lambda H'(\lambda) - H(\lambda) \log H(\lambda) = \mathrm{Ent}(f^2) \le 4 \frac{\lambda^2}{4} E_{\theta}\left( F e^{\lambda F} \right)= \lambda^2 H'(\lambda).
\]
Define $K(\lambda)= \frac{1}{\lambda}\log H(\lambda)$, with $K(0)= E_\theta(F)$. we get
\[
\begin{split}
K'(\lambda)= \frac{H'(\lambda)}{\lambda H(\lambda)} - \frac{\log H(\lambda)}{\lambda^2}.
\end{split}
\]
Thus
\[
\begin{split}
\lambda^2 H(\lambda) K'(\lambda) &= \mathrm{Ent}(f^2) \le \lambda^2 H'(\lambda).\\
\text{Hence}\quad K'(\lambda) &\le \frac{H'(\lambda)}{H(\lambda)}= \left( \lambda K(\lambda) \right)'= K(\lambda) + \lambda K'(\lambda)\\
\text{or}\quad \frac{K'(\lambda)}{K(\lambda)} &\le \frac{1}{1-\lambda}, \quad \text{or}\quad \left( \log K(\lambda) \right)' \le \frac{1}{1-\lambda}.
\end{split}
\]
Therefore
\[
\begin{split}
\log K(\lambda) &= \log K(0) + \int_0^\lambda (\log K(u))' du \le \log E_\theta(F) + \int_0^{\lambda} \frac{du}{1-u}\\
&= \log E_\theta(F) - \log(1-\lambda).
\end{split}
\]
Exponentiating both sides we get
\[
\begin{split}
K(\lambda) &\le (1-\lambda)^{-1}E_\theta(F)\\
\text{or}\quad H(\lambda) &\le \exp\left( \frac{\lambda}{1-\lambda} E_\theta(F) \right).
\end{split}
\]
Now, for any $r > 1$, we apply Chebyshev inequality to obtain
\[
\begin{split}
\mu^\theta\left( F \ge r E_\theta(F) \right) &=\mu^\theta\left( e^{\lambda F} \ge e^{\lambda r E_\theta(F)} \right)\le e^{- \lambda r E_\theta(F)} H(\lambda)\\
&\le \exp\left( \left( \frac{\lambda}{1-\lambda} -\lambda r \right) E_\theta F \right).
\end{split}
\]
We minimize the right side over $0 < \lambda < 1$ for a fixed $r$. That is, let
\[
\psi(\lambda)=\frac{\lambda}{1-\lambda} - \lambda r, \quad \text{then}\quad \psi'(\lambda)= \frac{1}{(1-\lambda)^2} - r.
\]
The minimum is attained at $\lambda=1-1/\sqrt{r}$ for which we get
\[
\frac{\lambda}{1-\lambda} - \lambda r= \sqrt{r}\left( 1-\frac{1}{\sqrt{r}} \right) - \left(1- \frac{1}{\sqrt{r}} \right)r=2\sqrt{r}-1-r=-(\sqrt{r}-1)^2.
\]
Thus
\[
\mu^\theta\left( F \ge r^2 E_\theta(F) \right) \le \exp\left( -(r-1)^2 E_\theta(F) \right), \quad \text{for any}\; r > 1.
\]
}
\begin{thm}\label{tauconcen}
The random variable $\chi=\max_i G_i$, where $G_i$'s are iid Gamma($\theta$), satisfies the following concentration estimate
\begin{equation}\label{eq:chicon}
P\left( \sqrt{\chi} > E(\sqrt{\chi}) + r \right) \le e^{-r^2}, \quad \text{for all}\;\; r > 0.
\end{equation}
\end{thm}
\begin{proof}[Proof of Theorem \ref{tauconcen}]
To prove \eqref{eq:chicon} we start by noting that Lemma \ref{expconcen} is satisfied by the family of $\mathbb{L}^k$-norms, $\{ F_{k}, \; k > 1 \}$, defined by
\[
F_k(x)= \left(\sum_{i=1}^n x_i^k\right)^{1/k}.
\]
This is because each $F_k$ is smooth (when every $x_i$ is positive) and
\begin{equation}\label{ineqcheck}
\begin{split}
\sum_{i=1}^n x_i \left(\partial_i F_{k}(x)\right)^2 &= \sum_{i=1}^n x_i \left[ \frac{x_i^{k-1}}{\left(\sum_{j=1}^n x_j^k \right)^{1-1/k}} \right]^2=\frac{\sum_{i=1}^n x_i^{2k-1}}{\left( \sum_{j=1}^n x_j^k \right)^{2-2/k}}.
\end{split}
\end{equation}
Since for any nonnegative $y_1, y_2, \ldots, y_n$ and any $\beta >1$ one has
\[
\sum_{i=1}^n y_i^{\beta} \le \left( \sum_{i=1}^n y_i \right)^\beta,
\]
applying it for $y_i= x_i^k$ and $\beta= 2-1/k$ we get
\[
\sum_{i=1}^n x_i^{2k-1} \le \left( \sum_{i=1}^n x_i^{k} \right)^{2-1/k}.
\]
Combining the above with \eqref{ineqcheck} we get
\[
\sum_{i=1}^n x_i \left(\partial_i F_{k}(x)\right)^2 \le \left(\sum_{i=1}^n x_i^k\right)^{1/k}=F_k(x).
\]
Thus $F_{k}$ satisfies condition \eqref{derivbnd}.
Since $F_k$ converges pointwise to $\max_i x_i$ as $k$ tends to infinity, by applying DCT, Lemma \ref{expconcen} is true for the function $\max_i G_i$. This proves \eqref{eq:chicon}.
\end{proof}
\comment{
To prove \eqref{eq:chicon2} we start by noting that (see, for example, \cite[p.~2716, Remark 1]{BW}) the product measure of iid Gamma$(\theta)$ distributions satisfies the following Poincar\'e inequality
\[
Var\left( f(G_1, \ldots, G_n) \right) \le 12\theta E\left( \abs{\nabla f}^2(G_1, \ldots, G_n) \right).
\]
One can take $f(x_1, \ldots, x_n)=\max_i x_i$, for which $\abs{\nabla f}^2$ is defined almost surely and is equal to one. The bound \eqref{eq:chicon2} is obtained by the Cauchy-Schwarz inequality.
The claim \eqref{eq:taucon} is obtained as a corollary by changing the variable from $\chi$ to $\tau$.
}
Our next step will be to prove estimate on the quantity $\sigma_0$ in \eqref{whatissigma0}. The process $\zeta(s)$ is non-Markovian and not distributed as $Q^{-n\theta}$. However, on an possibly enlarged sample space one can create a $Q^{-n\theta}$ process $\tilde \zeta$ such that the paths of $\zeta$ and $\tilde\zeta$ are indistinguishable until $\sigma_0$. This is possible by considering the SDE solved by $\zeta$:
\[
\zeta(t)= n - n \theta t + \int_0^t \sqrt{\zeta(s)} d W(s), \quad t < \sigma_0.
\]
To extend the process beyond $\sigma_0$, one concatenates an independent Brownian motion $\widetilde W$ and defines
\[
\beta(t) = \begin{cases}
W(t), &\quad t \le \sigma_0\\
W(t) + \widetilde W(t-\sigma_0), &\quad t > \sigma_0.
\end{cases}
\]
Then $\beta$ is a Brownian motion in the enlarged filtration. Since $Q^{-n\theta}$ admits a strong solution, the process
\begin{equation}\label{extendzeta}
\tilde\zeta(t) = n - n \theta t + 2\int_0^t \sqrt{\tilde\zeta(s)} d \widetilde W(s),\quad t < T_0,
\end{equation}
has law $Q^{-n\theta}$ and pathwise indistinguishable from $\zeta$ until time $\sigma_0$. Thus in the following discussion we will treat as if $\zeta$ itself is distributed as $Q^{-n\theta}$ keeping in mind the above construction.
\comment{
We also need the following observation about the process $\int_0^t ds/ \zeta(s)$, often called the \textit{Bessel clock}.
A similar result for the case of BESQ processes of positive dimensions has been proved by Yor and Zani in \cite{YZ}.
Consider the canonical space of continuous sample paths of nonnegative diffusions $C(\mathbb{R}^+, \mathbb{R}^+)$ along with the canonical right continuous filtration $\mcal{F}_t$ obtained from $\sigma\left\{ \omega(s), \; 0\le s \le t \right\}$.
\begin{lemma}\label{changeofmeasure}
For any $\theta > 0$ and $x > 0$, the probability measures $Q_x^{-\theta}$ and $Q_x^0$ satisfy the mutual absolute continuity condition
\begin{equation}\label{cofmeasure}
Q^{-\theta}_x\left( A \cap \left\{ T_0 > t \right\} \right)= Q^0_x\left(1\left\{A\cap \{ T_0 > t \}\right\}\left( \frac{\omega(t)}{x} \right)^{-\theta/4} \exp\left\{ -\frac{\theta(\theta+4)}{8} \int_0^t \frac{ds}{\omega(s)} \right\}\right)
\end{equation}
for all $A \in \mcal{F}_t$. Here
$T_0=\inf\left\{ t \ge 0:\; \omega(t)=0\right\}$.
\end{lemma}
\begin{proof}
This is an application of Girsanov's theorem. Suppose the process $Z$ has law $Q_x^0$, and consider the process
\[
M(t)= Z^{-\theta/4}(t\wedge T_0)H(t),\quad H(t)= \exp\left( -\frac{\theta(\theta+4)}{8}\int_0^{t\wedge T_0} \frac{ds}{Z(s)} \right).
\]
For $t < T_0$, by an application of It\^o's rule we get that
\[
\begin{split}
dM(t)&= H(t) d Z^{-\theta/4}(t) + Z^{-\theta/4}(t)dH(t)\\
&= H(t)\left[ -\frac{\theta}{4}Z^{-\theta/4-1}d Z(t) + \frac{\theta}{8}\left( \frac{\theta}{4}+1 \right)Z^{-\theta/4-2}(t)d\iprod{Z}(t) \right]\\
&-\frac{\theta(\theta+4)}{8}Z^{-\theta/4-1}(t) H(t) dt\\
&= H(t)\left[ -\frac{\theta}{2} Z^{-\theta/4-1/2}d\beta(t) \right] \\
&+ H(t)Z^{-\theta/4-1}(t)\left[ \frac{\theta}{2}\left( \frac{\theta}{4}+1 \right) - \frac{\theta(\theta+4)}{8}\right]dt\\
&= -\frac{\theta}{2} Z^{-(\theta+2)/4}(t)H(t)d\beta(t)=-\frac{\theta}{2} Z^{-1/2}(t)M(t)d\beta(t).
\end{split}
\]
Thus $M(t)$ is a local martingale.
For any $ x> \epsilon >0$, let $T_{\epsilon}=\inf\left\{ t\ge 0: \; Z(t)=\epsilon \right\}$. Then $Q_x^0(T_{\epsilon} < \infty)=1$, and $M(t\wedge T_{\epsilon})$ is a bounded martingale starting from $x^{-\theta/4}$. Thus, on the canonical space we define the change of measure by the recipe
\[
R_{x,\epsilon}(A)= Q_x^0\left( 1\{A\} x^{\theta/4}M(t\wedge T_{\epsilon}) \right), \quad \text{for all}\quad A \in \mcal{F}_t.
\]
Here, by an abuse of notation, we continue to refer to the martingale by $M$.
By Girsanov's Theorem (\cite[p.~ 327]{RY} ), it follows that under $R_{x, \epsilon}$, the process $\beta(t) = \int_0^t 2^{-1}\omega^{-1/2}(t)d\omega(t)$ (which is Brownian motion under $Q_x^0$) satisfies the SDE
\[
\begin{split}
d\beta(t) &= d\beta^*(t) + \frac{1}{M(t\wedge T_{\epsilon})}d\iprod{M(t\wedge T_{\epsilon}), \beta(t)}\\
&= d\beta^*(t) -\frac{\theta}{2M(t)} \omega^{-1/2}(t) M(t) d(t\wedge T_{\epsilon})\\
&=d\beta^*(t) - \frac{\theta}{2\omega^{1/2}}d(t\wedge T_{\epsilon}).
\end{split}
\]
In other words, under the changed measure $R_{x,\epsilon}$, the stopped process $\omega(t\wedge T_{\epsilon})$ satisfies the SDE
\[
d\omega(t \wedge T_{\epsilon}) = -\theta d(t\wedge T_{\epsilon}) + 2\sqrt{\omega(t)} d \beta^*(t\wedge T_{\epsilon}),
\]
and hence has the law of a stopped $Q_x^{-\theta}$.
In particular, we get that for any $A \in \mcal{F}_t$,
\[
Q_x^{-\theta}\left( A\cap\{ t < T_{\epsilon} \}\right)= x^{\theta/4} Q^0_x\left( 1\{A\cap\{ t < T_{\epsilon} \} M(t) \right).
\]
We let $\epsilon$ go to zero and apply the Monotone Convergence Theorem to get \eqref{cofmeasure}.
\end{proof}
\begin{thm}\label{thm:formlaplace}
Let $Z$ be distributed as $Q_x^{-\theta}$ and let $C_t$ denote the \textit{Bessel clock} integral
\[
C_t = \int_0^{t\wedge T_0} \frac{ds}{Z(s)}.
\]
Then the formula for the Laplace transform for the law of $C_t$ is given by
\begin{equation}\label{formlaplace}
Q_x^{-\theta}\left( e^{-\phi C_t} \right)= \left( \frac{x}{2t} \right)^{\kappa/2+1-\alpha} E\left( \frac{\Gamma(\alpha+Y+1)}{\Gamma(\kappa/2+Y+2)} \right),
\end{equation}
Here, $Y$ denotes a Poisson random variable with mean $x/2t$, and
\[
\kappa = -2 + \sqrt{(\theta+2)^2 + 8 \phi}, \quad \alpha = (\kappa - \theta)/4.
\]
\end{thm}
\begin{proof} For BESQ processes of any dimension $-\theta \in \mathbb{R}$, define the index as
\[
\nu = -\theta/2 -1.
\]
It will be notationally convenient to use the index of the BESQ processes as the parameter.
Now, let $\phi$ and $\kappa$ as in the statement of the Theorem, and note that $\kappa > 0$. Moreover
\[
\begin{split}
\frac{\kappa(\kappa + 4)}{8}&= \frac{\left(-2 + \sqrt{(\theta+2)^2 + 8 \phi}\right)\left( 2 + \sqrt{(\theta+2)^2 + 8 \phi}\right)}{8}\\
&=\frac{1}{8}\left[ -4 + (\theta+2)^2 + 8\phi \right]=\frac{\theta(\theta+4)}{8}+\phi.
\end{split}
\]
By using the change of measure \eqref{cofmeasure} we get
\begin{equation}\label{momenteq}
\begin{split}
Q_x^{-\theta}\left( e^{-\phi C_t}1\left\{ T_0 > t \right\}\right)&= Q_x^0\left[ 1\left\{ T_0 > t \right\} \left( \frac{\omega(t)}{x} \right)^{-\theta/4} \exp\left\{ \left(-\phi -\frac{\theta(\theta+4)}{8}\right)C_t \right\}\right]\\
&=Q_x^0\left[ 1\left\{ T_0 > t \right\} \left( \frac{\omega(t)}{x} \right)^{-\theta/4} \exp\left\{ -\frac{\kappa(\kappa+4)}{8}C_t \right\}\right]\\
&= Q_x^{-\kappa}\left( 1\left\{ T_0 > t \right\} \left( \frac{\omega(t)}{x} \right)^{\alpha} \right).
\end{split}
\end{equation}
Here
\[
\alpha=\kappa/4-\theta/4= -\frac{(\theta+2)}{4}+ \frac{1}{4}\sqrt{(\theta+2)^2 + 8 \phi}=\frac{\nu}{2} + \frac{1}{2}\sqrt{\nu^2+2\phi} > 0.
\]
We now compute the moment above in \eqref{momenteq}. Note that by scaling
\[
\begin{split}
x^{-\alpha}Q_x^{-\kappa}\left( 1\left\{ T_0 > t \right\} \omega(t)^{\alpha} \right)=x^{-\alpha}t^{\alpha}Q_{x/t}^{-\kappa}\left( 1\left\{ T_0 > 1 \right\} \omega(1)^{\alpha} \right).
\end{split}
\]
We may define $C_{t}\equiv \infty$ for all $t \ge T_0$. Hence, combining all the previous steps we get
\[
Q_x^{-\theta}\left( e^{-\phi C_t} \right)= \left( \frac{t}{x} \right)^\alpha \int_0^\infty y^{\alpha}q_{1}^{-\kappa}(x/t,y)dy.
\]
This proves the Laplace transform formula \eqref{formlaplace}.
We can evaluate the above moment using the explicit description of the transition density given in \eqref{tranden}. That is
\[
\begin{split}
\int_0^\infty y^{\alpha}q_{1}^{-\kappa}(x/t,y)dy&= \sum_{j=0}^\infty g_{\kappa + 4 + 2j}(x/t) \frac{2^{-j}}{j!} \int_0^{\infty} y^{\alpha+j} e^{-y/2}dy\\
&= \sum_{j=0}^\infty g_{\kappa + 4 + 2j}(x/t) \frac{2^{-j}}{j!} 2^{\alpha+j+1}\Gamma(\alpha+j+1)\\
&= 2^{\alpha+1}\sum_{j=0}^\infty g_{\kappa + 4 + 2j}(x/t) \frac{\Gamma(\alpha+j+1)}{j!}\\
&= 2^{\alpha+1}\sum_{j=0}^\infty \frac{2^{-\kappa/2-2-j} (x/t)^{\kappa/2+1+j}}{\Gamma(\kappa/2+2+j)} \frac{\Gamma(\alpha+j+1)}{j!} e^{-x/2t}\\
&= 2^\alpha \left( \frac{x}{2t} \right)^{\kappa/2+1} e^{-x/2t}\sum_{j=0}^\infty \frac{\Gamma(\alpha+j+1)}{\Gamma(\kappa/2+j+2)} \frac{1}{j!}\left( \frac{x}{2t} \right)^j
\end{split}
\]
Thus, if $Y$ denotes a Poisson random variable with mean $x/2t$, we get
\begin{equation}\label{ident1}
Q_x^{-\theta}\left( e^{-\phi C_t} \right)= \left( \frac{x}{2t} \right)^{\kappa/2+1-\alpha} E\left( \frac{\Gamma(\alpha+Y+1)}{\Gamma(\kappa/2+Y+2)} \right).
\end{equation}
This proves formula the result.
\end{proof}
Now we make the following choice of parameters for computational convenience
\[
\alpha=1, \quad \kappa= \theta+4, \quad \phi=\frac{1}{8}\left[ (\kappa+2)^2 - (\theta+2)^2 \right]= \theta+4.
\]
In this case
\begin{equation}\label{poissonexpec}
\begin{split}
E&\left( \frac{\Gamma(\alpha+Y+1)}{\Gamma(\kappa/2+Y+2)} \right)= e^{-x/2t}\sum_{j=0}^\infty \frac{(j+1)!}{\Gamma(\kappa/2 + j + 2)} \frac{1}{j!}\left( \frac{x}{2t} \right)^j\\
&= e^{-x/2t} \sum_{j=1}^\infty \frac{j}{\Gamma(\kappa/2 + j + 2)} \left( \frac{x}{2t} \right)^j + e^{-x/2t} \sum_{j=0}^\infty \frac{1}{\Gamma(\kappa/2 + j + 2)} \left( \frac{x}{2t} \right)^j.
\end{split}
\end{equation}
Let $\gamma^*(a,x)$ denote the incomplete Gamma function (\cite[p.~260]{AS})
\begin{equation}\label{whatisgammast}
\gamma^*(a,x)= \frac{x^{-a}}{\Gamma(a)}\int_0^x e^{-t} t^{a-1}dt.
\end{equation}
Then it is known (\cite[p.~262]{AS}) that $\gamma^*(a,x)$ admits the follows series expansion
\[
\gamma^*(a,x)= e^{-x} \sum_{n=0}^\infty \frac{x^n}{\Gamma(a+n+1)}.
\]
Substituting this value in \eqref{poissonexpec} we get
\begin{equation}\label{ident2}
\begin{split}
E\left( \frac{\Gamma(\alpha+Y+1)}{\Gamma(\kappa/2+Y+2)}\right)&= xe^{-x/2t} \frac{d}{dx}\left( e^{x/2t}\gamma^*(\kappa/2+1, x/2t) \right) + \gamma^*(\kappa/2+1,x/2t)\\
&= \left( 1 + \frac{x}{2t}\right)\gamma^*(\kappa/2+1, x/2t) + x \frac{d}{dx} \gamma^*(\kappa/2+1, x/2t).
\end{split}
\end{equation}
Now, from \eqref{whatisgammast}, we get
\begin{equation}\label{ident3}
\frac{d}{dx} \gamma^*(a,x)= \frac{x^{-1}e^{-x}}{\Gamma(a)} -a \frac{x^{-a-1}}{\Gamma(a)} \int_0^x e^{-t} t^{a-1}dt.
\end{equation}
Combining \eqref{ident1}, \eqref{ident2}, and \eqref{ident3} we get
\[
\begin{split}
Q_x^{-\theta}\left( e^{-\phi C_t} \right)&= \left( \frac{x}{2t} \right)^{\kappa/2+1-\alpha} \left[ \left( 1 + \frac{x}{2t}\right)\gamma^*(\kappa/2+1, x/2t) + x \frac{d}{dx} \gamma^*(\kappa/2+1, x/2t) \right]\\
&=\left( \frac{x}{2t} \right)^{\kappa/2}\left[ \left( 1 + \frac{x}{2t}\right)\frac{(x/2t)^{-\kappa/2-1}}{\Gamma(\kappa/2+1)}\int_0^{x/2t} e^{-u}u^{\kappa/2}du + \frac{e^{-x/2t}}{\Gamma(\kappa/2+1)} \right]\\
&-\left( \frac{x}{2t} \right)^{\kappa/2+1}\left[(\kappa/2+1)\frac{(x/2t)^{-\kappa/2-2}}{\Gamma(\kappa/2+1)}\int_0^{x/2t} e^{-u}u^{\kappa/2}du \right] \\
&= \left( 1+ \frac{2t}{x} \right) P\left( G \le x/2t \right) + \frac{1}{\Gamma(\kappa/2 + 1)}\left( \frac{x}{2t} \right)^{\kappa/2}e^{-x/2t}\\
&- (\kappa/2+1)\frac{2t}{x} P\left( G \le x/2t \right).
\end{split}
\]
Here $G$ is obviously distributed as Gamma$(\kappa/2+1)$.
As a consequence of the previous Theorem we get the following corollary.
\begin{cor}\label{cor:lbnd}
Consider $C_t$ as in Theorem \ref{thm:formlaplace} and a family of pairs $(x,t)$ such that $\{x/t\}$ is contained in $(0,1)$. Let $H(u)$ denote the cumulative distribution function of a Gamma random variable with parameter $(\theta/2 + 3)$. Then
\[
\begin{split}
Q_x^{-\theta}&\left( C_t < u \right) \le C_\theta e^{(\theta+4)u}\left( \frac{x}{2t} \right)^{2+ \theta/2},
\end{split}
\]
for some constant $C_\theta$ depending on $\theta$.
\end{cor}
\begin{proof} In the equality \eqref{formlaplace} make the following choice of parameters
\[
\alpha=1, \quad \kappa= \theta+4, \quad \phi=\frac{1}{8}\left[ (\kappa+2)^2 - (\theta+2)^2 \right]= \theta+4.
\]
Thus
\[
Q_x^{-\theta}\left( e^{-(\theta+4) C_t} \right)=\left( \frac{x}{2t} \right)^{2+ \theta/2} E\left( \frac{\Gamma(Y+2)}{\Gamma(Y+4+\theta/2)} \right).
\]
Since $x/t$ is bounded by one, the expectation of the right side is bounded by a nonzero constant. This proves the corollary by Markov's inequality.
\end{proof}
\begin{lemma}
Suppose the process $\zeta$ is distributed as $Q_1^{-2n\delta}$ for some $\delta >0$. Then, for a sequence $\{a_n\}$ of positive constants such that $\lim_{n\rightarrow \infty} na_n=0$, we get
\[
Q_1^{-2n\delta} \left( \frac{a_n}{n(1+\epsilon)} < \int_0^{a_n} \frac{ds}{\zeta(s)} < \frac{a_n}{n(1-\epsilon)} \right) <
\]
\end{lemma}
\begin{proof} Consider the event $A_n=\{ 1-\epsilon < \zeta_s < 1+\epsilon, \; \forall \; 0\le s \le a_n \}$. By Lemma \ref{changeofmeasure} we get
\[
Q_1^{-2n\delta} \left( A_n^c \right) \ge (1+\epsilon)^{n\delta/2} \exp\left\{-\frac{n\delta(n\delta + 2)a_n}{2(1-\epsilon)} \right\}Q^0_1(A_n).
\]
Let $\sigma_\epsilon$ denote the hitting time of either of the boundaries $1\pm \epsilon$ for some $\epsilon \approx 0$. Since $\zeta$ is a martingale under $Q^0$, we get
\[
\begin{split}
Q_1^{0}\left( \sigma_\epsilon < a_n \right) &= Q_1^{0} \left( \sup_{0\le s\le a_n} \abs{\zeta_s - 1} > \epsilon \right)< c\epsilon^{-2} Q_1^{0} \iprod{\zeta}_{a_n}.
\end{split}
\]
The constant $c$ is universal and comes from the Burkholder-Davis-Gundy inequality.
However,
\[
Q_1^0 \iprod{\zeta}_{a_n} = 2 \int_0^{a_n} Q_1^0\left( \zeta_s\right) ds = 2 a_n.
\]
Hence
\[
Q_1^{-2n\delta} \left( A_n \right) \ge (1-2a_n) (1+\epsilon)^{n\delta/2} \exp\left\{-\frac{n\delta(n\delta + 2)a_n}{2(1-\epsilon)} \right\}.
\]
\end{proof}
}
\begin{thm}\label{thm:exittime}
Let $\mu$ be distributed as an $n$-dimensional NWF$(\delta, \delta, \ldots, \delta)$ starting from the point $(1/n, 1/n,\ldots, 1/n)$. Let $\sigma_0$ be the first time that any of the coordinates of $\mu$ hit zero. Let
\[
a_n = E \max_{1\le i \le n} \sqrt{G_i}, \qquad G_i \stackrel{\text{iid}}{\sim} \text{Gamma}(\delta+1).
\]
Then, $a_n = \Theta(\sqrt{\log n})$, $\sigma_0$ has the law given by \eqref{whatissigma0} where $\zeta$ is distributed as $Q^{-2n\delta}_1$ and $\tau$ is a random time.
Moreover, for any $r > 0$, we get
\[
\begin{split}
P\left( \frac{1}{n (a_n + r)} \le \sqrt{2\tau} \le \frac{1}{n(a_n+r)}\right) \ge 1 - 2 e^{-r^2}.
\end{split}
\]
\end{thm}
\begin{rmk}
It is impossible to provide a simple description of the exact distribution of $\sigma_0$ due to the distributional dependence of $\zeta$ and $\tau$. The above theorem shows that $\tau$ is about a constant and one can compare the distribution of $\sigma_0$ with that of $4\int_0^{\cdot} du/\zeta(u)$, where the upper limit of the integral is a constant. Limiting large deviation behavior of such integrals is possible to derive by methods as in \cite{YZ}.
\end{rmk}
\begin{proof}[Proof of Theorem \ref{thm:exittime}] The proof is obvious from Lemma \ref{tauconcen} and expression \eqref{whatistau}.
\end{proof}
\section*{Appendix: \\ Proofs of properties of BESQ processes}
\begin{proof}[Proof of Lemma \ref{scale}] We use exercise 3.20 in \cite[p.~311]{RY}. The scale function for $Q^{\theta}$ for $\theta \ge 0$ is well-known to be $x^{-\theta/2+1}$ (see \cite[p.~443]{RY}). Nearly identical calculations lead to the case when $\theta$ is replaced by $-\theta$ and we obtain the scale function $s(x)= x^{\theta/2+1}$.
The speed measure is the measure with the density
\[
m'(x)=\frac{2}{s'(x)4x}=\frac{1}{2(\theta/2+1)}x^{-\theta/2-1}.
\]
We now Feller's criterion to check if the origin is an entry and/or exit point (see \cite[p.~108]{itomckean}). Note that
\begin{equation}
\begin{split}
m(\xi,1/2)&=\frac{1}{2(\theta/2+1)}\int_{\xi}^{1/2}x^{-\theta/2-1}dx=\frac{1}{\theta(\theta/2+1)}\left( \xi^{-\theta/2} - 2^{\theta/2} \right)\\
m(0,\xi]&=\infty, \quad \text{for all positive} \; \xi.
\end{split}
\end{equation}
Thus
\[
\int_0^{1/2} m(\xi,1/2] s(d\xi) < \infty, \quad \text{and}\quad \int_0^{1/2} m(0,\xi] s(d\xi)=\infty.
\]
This proves that the origin is an exit and not an entry.
Finally, to obtain part (ii) we apply Girsanov's theorem (\cite[p.~327]{RY}). Let $X$ satisfy the SDE $d X(t)= -\theta dt + 2\sqrt{X(t)} d\beta(t)$, then, we take $D(t)=X^{\theta/2+1}(t)$ (without the normalization, for simplicity) and apply Girsanov. Under the changed measure, there is a standard Brownian motion $\beta^*$ such that
\[
\begin{split}
\beta(t) &= \beta^*(t) + \int_0^t X^{-\theta/2-1}(s) d\iprod{\beta, D}_s\\
&=\beta^*(t) + \int_0^t X^{-\theta/2-1}(s) \left( \theta + 2 \right) X^{\theta/2+1/2}(s) ds=\beta^*(t) + \left( \theta + 2 \right) \int_0^t X^{-1/2}(s) ds.
\end{split}
\]
Thus, under the changed measure
\[
\begin{split}
dX(t) &= -\theta dt + 2 X^{1/2}(t)d\beta(t) = -\theta dt + 2(\theta +2)dt + d\beta^*(t) = (\theta+4)dt + d \beta^*(t).
\end{split}
\]
The interpretation as the conditional distribution is classical (see \cite{besselbridge}).
\end{proof}
\begin{proof}[Proof of Lemma \ref{logct}]
For the assertion it is enough to take $t=1$. Note that, under $Q_0^{\theta}$, the coordinate process satisfies time-inversion, i.e., the process $\{ t^2 Z(1/t), \quad t\ge 0\}$ has law $Q_0^\theta$. Thus, for $0<\epsilon < 1$, if we define
\[
U_{\epsilon}= \int_{\epsilon}^1 \frac{du}{Z(u)}= \int_{1}^{1/\epsilon} \frac{dt}{t^2 Z(1/t)},
\]
then $U_\epsilon$ has the same law as $C_{1/\epsilon} - C_1 = \int_{1}^{1/\epsilon} du/Z(u)$.
Thus, by \cite[Thm.~1.1]{YZ}, we get $\lim_{\epsilon \rightarrow 0} U_{\epsilon}/\log(1/\epsilon) = (\theta - 2)^{-1}$ almost surely.
\end{proof}
\section*{Acknowledgement}
\noindent I thank David Aldous, Zhen-Qing Chen, Michel Ledoux, and Jon Wellner for very useful discussions. I thank the anonymous referee for a thorough review which led to a significant improvement of the article.
\bibliographystyle{alpha}
|
\section{Introduction
The formulation of the cosmological principle~\cite{Einstein:1917ce} coincides with the birth of modern scientific cosmology. Over the last decades, homogeneity and isotropy of the Universe have been tested to increasing degrees of accuracy. In particular, the radiation of the Cosmic Microwave Background (CMB) is one of the most sensitive probes to test the isotropy of the Universe. The observed temperature of the CMB is observed to be uniform at zeroth order (once the dipole is subtracted): this result led to the formulation of the theory of primordial inflation. Moreover, approximate statistical isotropy appears to hold even for the small fluctuations of the CMB temperature.
In spite of these striking results, several analyses of the CMB fluctuation maps, starting from~\cite{Eriksen:2003db,Hansen:2004vq} (see, {\em e.g.}, \cite{Hanson:2009gu} for a more recent analysis), have shown the existence of anomalies associated to some degree of breaking of statistical isotropy. Even though a clear consensus on the subject is still absent ({\em e.g.}, see \cite{Bennett:2010jb} for an up to date discussion) intensive work has been done in the past also as a response to the theoretical challenge of naturally generating a preferred directions in the sky. The models proposed usually rely either on early-Universe or on late-Universe mechanisms. While the former lead to an {\em intrinsic} anisotropy in the primordial spectrum of metric perturbations, the latter produce anisotropy in the {\em observed} CMB via anisotropic Sachs-Wolfe effect. Early-Universe mechanisms include~\cite{Ackerman:2007nb} (see however~\cite{Himmetoglu:2008hx}) and~\cite{Watanabe:2009ct}, that rely on vector fields or~\cite{Boehmer:2007ut}, that rely on spinor fields, as well as~\cite{Donoghue:2007ze,Erickcek:2008sm}, that rely on primordial gradients. Late-Universe mechanisms can invoke magnetic fields~\cite{Campanelli:2007qn} or anisotropies in the dark energy equation of state~\cite{Rodrigues:2007ny,Koivisto:2008ig,Battye:2009ze}. Despite those many attempts, it is fair to say that breaking statistical isotropy usually requires strong and aesthetically unappealing assumptions.
In the present work we show a (relatively) natural mechanism that can give rise to a preferred direction in the CMB sky. This mechanism is a hybrid of early and a late Universe ones and does not rely on vector fields or on unusual properties of the dark energy sector, but rather on a simple model of scalar dark energy, characterized by a non-trivial, tachyonic ($V''(\phi)<0$) potential.
Models of tachyonic quintessence frequently appear in the literature. For instance, if the quintessential scalar is described by a pseudo-Nambu-Goldstone Boson (pNGB)~\cite{Frieman:1995pm}, then half of the extrema of the potential $V(\phi)\propto 1+\cos(\phi/f)$ are tachyonic. Another scenario of tachyonic quintessence was proposed in~\cite{Kallosh:2002gf}, where the magnitude of the tachyonic mass is related to the height of the potential at its maximum.
Our main assumption is that a tachyonic field $\phi$ has a small primordial gradient. As the Hubble parameter drops below $\sqrt{\left|V''(\phi)\right|}$, the initial gradient is amplified by the ``pull'' of its tachyonic potential, effectively converting a small primordial isocurvature perturbation into a larger curvature mode. In first approximation, this leads to an ``ellipsoidal'' Universe~\cite{Campanelli:2007qn} which could explain the low quadrupole CMB amplitude observed in both COBE and WMAP data~\cite{Contaldi:2003zv}, if the preferred direction in the quintessence correlated with that of the primordial quadrupole.
As we will see, our mechanism will be at work for a tachyonic mass a few times larger than $\Lambda/M_\mathrm{Pl}^2$. This is, in particular, the case for a pNGB with value of the axion constant $f$ slightly sub-Planckian.
The primordial gradient in $\phi$ can be a remnant of the epoch that preceded the last bout of inflation, provided that such last bout were sufficiently short (this argument will be discussed in greater detail in section~\ref{gradient}). In this respect, our scenario could shed some light on pre-inflationary dynamics, similarly to the situations studied, {\em e.g.}, in~\cite{Chang:2007eq} and especially~\cite{Donoghue:2007ze}. In our case, a constant gradient can be the dominant remnant of a more general inhomogeneous initial value of the tachyonic field, {\em i.e.}, we can expand $\phi\approx \kappa_0+\kappa_1\,z_p+\kappa_2\,z^2_p+\dots$ -- for sake of simplicity, we will only consider powers of the $z$ coordinate, though the argument for the dominance of the linear term would still hold, were more generic terms considered -- at $N_\mathrm{i}\equiv N_{\mathrm{obs}}+\delta N$ e-foldings before the end of inflation ($N_{\mathrm{obs}}$ corresponds to the time when the current cosmological scales left the horizon and $z_p$ to the physical distance along the $z$ direction). We assume that each term $\kappa_k\,z_p^k$ had given an equal contribution to the energy in $\phi$ at that moment. Today, the scales, that left the horizon $N_{\mathrm{obs}}+\delta N$ e-foldings before the end of inflation, are still outside the horizon by a factor of ${\mathrm{e}}^{\delta N}$, {\em i.e.}, they correspond to physical distances of the order of ${\mathrm{e}}^{\delta N}\,H_0^{-1}$. Assuming no (other) significant evolution in $\phi$~\footnote{The evolution in $\phi$ occurs only at sub-horizon scales and in the late Universe, when the Hubble parameter is smaller than the (absolute value of the tachyonic) mass of quintessence, and it is not sufficient to invalidate this argument.}, the assumption that each term $\kappa_k\,z_p^k$ be of the same order implies that $\kappa_k\approx {\mathrm{e}}^{-k\,\delta N}\,H_0^k$. As a consequence, inside the horizon $z_p\approx H_0^{-1}$, each term $\kappa_k\,z_p^k$ contributes like ${\mathrm{e}}^{-k\,\delta N}$ to the energy in $\phi$. This implies that the terms with the lowest powers in $z_p$ will give the largest contributions to the metric perturbations at sub-horizon scales. For this reason, our analysis will be focused on the approximation where, in the early Universe, $\phi=\kappa_1\,z$ with $\phi(z=0)$ being set equal to zero by an appropriate choice of the origin of the $z$ coordinate.\\
The paper is organized as follows. In section~\ref{metric} we set up our system and solve the corresponding Einstein equations. Sections~\ref{rs} and~\ref{cmb} contain the effect of our anisotropic metric on the redshift and in particular on the CMB quadrupole anisotropy. Section~\ref{gradient} contains a discussion of the magnitude of the primordial gradient responsible for the anisotropy. Finally, in the concluding section we will discuss how our mechanism could also lead to alignments of the higher multipoles.
\section{The metric}\label{metric
We consider the cylindrically symmetric cosmological metric
\begin{equation}
\mathrm{d} s^2=-\mathrm{d} t^2+a(t,\,z)^2\,\left(\mathrm{d} x^2+\mathrm{d} y^2\right)+b(t,\,z)^2\,\mathrm{d} z^2\,\,.
\end{equation}
The non-vanishing components of the Einstein tensor are
\begin{eqnarray}
G_{tt}&=&\frac{2}{b^2}\,\frac{a'}{a}\,\frac{b'}{b}-\frac{1}{b^2}\,\left(\frac{a'{}^2}{a^2}+2\,\frac{a''}{a}\right)+\frac{\dot{a}^2}{a^2}+2\,\frac{\dot{a}}{a}\,\frac{\dot{b}}{b},\nonumber\\
G_{xx}&=&G_{yy}=a^2\,\left(-\frac{1}{b^2}\,\frac{a'}{a}\,\frac{b'}{b}+\frac{1}{b^2}\frac{a''}{a}-\frac{\dot{a}}{a}\frac{\dot{b}}{b}-\frac{\ddot{a}}{a}-\frac{\ddot{b}}{b}\right)\,\,,\nonumber\\
G_{tz}&=&2\,\frac{a'}{a}\frac{\dot{b}}{b}-2\frac{\dot{a}'}{a}\,\,,\nonumber\\
G_{zz}&=&b^2\left(\frac{1}{b^2}\frac{a'{}^2}{a^2}-\frac{\dot{a}^2}{a^2}-2\frac{\ddot{a}}{a}\right)\,\,,
\end{eqnarray}
where a dot denotes a derivative with respect to $t$ and a prime a derivative with respect to $z$.
We assume the presence of three contributions to the stress-energy tensor: pressure-less dust with energy density $\rho$, a cosmological constant $\Lambda$ and a scalar field $\phi$ with canonically normalized kinetic term and potential $V(\phi)$ -- in principle one can always consider $\Lambda$ to be a part of $V(\phi)$ however we prefer to keep these two components separate as we will keep separate the cosmological constant problem and the breaking of statistical isotropy.
The components of the stress energy tensor of $\phi$ read
\begin{eqnarray}
T_{tt}&=&\frac{1}{2}\,\dot{\phi}^2+\frac{1}{2\,b^2}\,\phi'{}^2+V(\phi)+\Lambda\,\,\nonumber\\
T_{xx}&=&T_{yy}=a^2\left[\frac{1}{2}\dot{\phi}^2-\frac{1}{2\,b^2}\,\phi'{}^2-V(\phi)-\Lambda\right]\,\,,\nonumber\\
T_{zt}&=&\dot\phi\,\phi'\,\,,\nonumber\\
T_{zz}&=&b^2\left[\frac{1}{2}\dot{\phi}^2+\frac{1}{2\,b^2}\,\phi'{}^2-V(\phi)-\Lambda\right]\,\,,
\end{eqnarray}
and throughout the paper we will consider $\phi$ to be a tachyon with potential $V(\phi)=-m^2\,\phi^2/2$. The stress energy tensor for dust has the general form $T_{\mu\nu}=\rho\,u_\mu\,u_\nu$, with $u^\mu\,\nabla_\mu u^\nu=0$.
Let us now solve the above equations by assuming that the non-isotropic, non-homogeneous scalar field is the source of a perturbation around the Friedmann-Robertson-Walker metric of a $\Lambda\mathrm{CDM}$ Universe. As a consequence, we write $a(t,\,z)=a_0(t)\,[1+\delta a(t,\,z)]$ and $b(t,\,z)=a_0(t)\,[1+\delta b(t,\,z)]$, where $\delta a\ll 1$ and $\delta b\ll 1$; the Friedmann equation for the background $\phi=0$ has exact solution
\begin{equation}\label{scalefactor}
a_0(t)=\left(\frac{1-\Omega_\Lambda}{\Omega_\Lambda}\right)^{1/3}\,\sinh^{2/3}\left(\frac{3}{2}\,\sqrt{\Omega_\Lambda}\,H_0\,t\right)\,\,,
\end{equation}
where $H_0$ is the current value of the Hubble parameter and $\Omega_\Lambda\sim0.3$ from observation.
We can assume that the dust velocity field $u^\mu$ is equal to $(1+\delta u^0,\, 0,\,0,\,\delta u^z)$ with $\delta u^0$ and $\delta u^z$ being first order quantities. Thus the geodesic equation for $u^\mu$ reduces to $\delta\dot{u}^0=0$ and $\left(a_0^2\,\delta u^z\right)^.=0$. Choosing $\delta u^z=0$ as $t\rightarrow 0$, we obtain, up to second order in the perturbations, $u^\mu=(1,\,0,\,0,\,0)$. Then conservation of the stress-energy tensor for dust implies that $\rho=\rho_0/(a^2\,b)$, where $\rho_0$ is the dust density at a fixed time.
Turning on the scalar field $\phi$, we assume that, as described in the introduction (see also~\cite{Turner:1991dn,Langlois:1995ca,Erickcek:2008sm}), the lowest terms in the expansion in powers of the coordinate $z$ give the largest contributions to the anisotropy. Then, without loss of generality, we can set $\phi(z=0)=0$ (remember that $\phi=0$ corresponds to the maximum of the tachyonic potential), so that
\begin{equation}
\phi(t,\,z)=\kappa_1(t)\,z+{\cal {O}}(z^2)\,\,.
\end{equation}
At leading order, the Klein-Gordon equation $\ddot\phi+3H\dot\phi-\phi''/a_0^2-m^2\,\phi=0$ decomposes as
\begin{eqnarray}
\ddot\kappa_1+3\frac{\dot a_0}{a_0}\,\dot\kappa_1-m^2\,\kappa_1=0\,\,.\label{k1}
\end{eqnarray}
We then expand $\delta a(t,\,z)=\delta a_0(t)+\delta a_2(t)\,z^2+{\cal {O}}(z^3)$ and $\delta b(t,\,z)=\delta b_0(t)+\delta b_2(t)\,z^2+{\cal {O}}(z^3)$ (it is easy to see that, since at this level of approximation $\phi$ is odd in $z$, $\delta a$ and $\delta b$ must be even in $z$). The $(tz)$ Einstein equation -- at first order $-2\,\delta \dot{a}'=\phi'\,\dot\phi/M_\mathrm{Pl}^2$ -- gives rise to
\begin{eqnarray}
\delta\dot{a}_2&=&-\frac{\kappa_1\,\dot\kappa_1}{4\,M_\mathrm{Pl}^2}\,\,.\label{deltaa2}
\end{eqnarray}
The $(zz)$ Einstein equation reads
\begin{equation}
-2\,\delta\ddot{a}-6\,\frac{\dot{a}_0}{a_0}\,\delta\dot a=\frac{1}{M_\mathrm{Pl}^2}\left(\frac{1}{2}\dot\phi^2+\frac{1}{2\,a_0^2}\phi'{}^2+\frac{m^2}{2}\phi^2\right)\,\,.
\end{equation}
The only non-trivial part of this equation is the one independent on $z$, as the rest is redundant with respect to the two previous equations:
\begin{equation}\label{deltaa0}
-2\,\delta\ddot{a}_0-6\,\frac{\dot{a}_0}{a_0}\,\delta\dot a_0=\frac{\kappa_1^2}{2\,a_0^2\,M_\mathrm{Pl}^2}\,\,.
\end{equation}
The $(tt)$ Einstein equation reads, at first order in $\delta a$ and $\delta b$ and using the background equations,
\begin{eqnarray}
&&-2\frac{\delta a''}{a_0^2}+2\,\frac{\dot a_0}{a_0}\,\left(2\,\delta \dot a+\delta \dot b\right)=\\
&&=\frac{1}{M_\mathrm{Pl}^2}\left[\frac{1}{2}\dot\phi^2+\frac{1}{2\,a_0^2}\phi'{}^2-\frac{m^2}{2}\phi^2-\frac{\rho_0}{a_0^3}\left(2\,\delta a+\delta b\right)\right]\,\,,\nonumber
\end{eqnarray}
where the only term under control in the $z^n$ expansion is the one in $z^0$ (since $\delta a$ is determined with $\mathcal{O}(z^3)$ accuracy, the presence of the term $\delta a''$ constrains the accuracy of this equation to be $\mathcal{O}(z)$ hence, at this level of approximation, $\delta b$ can be determined only in its $z^0$ component). Therefore the only equation we obtain from the $(tt)$ component is
\begin{eqnarray}
\label{equdeltab0}
-4\frac{\delta a_2}{a_0^2}+2\,\frac{\dot a_0}{a_0}\,\delta \dot \psi_0&=&\frac{1}{M_\mathrm{Pl}^2}\left(\frac{\kappa_1^2}{2\,a_0^2}-\frac{\rho_0}{a_0^3}\,\delta \psi_0\right)\,\,,
\end{eqnarray}
where $\delta\psi_0\equiv 2\,\delta a_0+\delta b_0$.\\
The explicit expressions of $\kappa_1(t)$, $\delta a_0(t)$ and $\delta b_0(t)$ are the following. The solution of eq.~\eqref{k1}, using the expression~\eqref{scalefactor} for $a_0(t)$, reads
\begin{equation}
\kappa_1(t)=\frac{\bar\kappa_1\,H_0\,M_\mathrm{Pl}}{\sqrt{1+\mu^2}}\frac{\sinh(\sqrt{1+\mu^2}\tau)}{\sinh\tau}
\end{equation}
where we defined the dimensionless quantities $\tau\equiv\frac{3}{2}\sqrt{\Omega_\Lambda}H_0\,t$ and $\mu^2\equiv4\,m^2/(9\,\Omega_\Lambda H_0^2)$, and where the integration constant $\bar\kappa_1$ is defined so that $\kappa_1(0)=\bar\kappa_1\,H_0\,M_\mathrm{Pl}$.
It is straightforward to solve for $\delta a_2$ from eq.~\eqref{deltaa2} using the found solution for $\phi(t,\,z)$ at the desired order:
\begin{eqnarray}
\delta a_2(t)&=&-\frac{\kappa_1(t)^2-\bar\kappa_1^2H_0^2M_\mathrm{Pl}^2}{8\,M_\mathrm{Pl}^2}\,\,,
\end{eqnarray}
where we have imposed $\delta a_2(t\rightarrow 0)=0$.
$\delta a_0$ can be computed by integrating eq.~\eqref{deltaa0}:
\begin{eqnarray}
\delta a_0(t)=-\frac{1}{9\,\Omega_\Lambda H_0^2M_\mathrm{Pl}^2}\left(\frac{\Omega_\Lambda}{1-\Omega_\Lambda}\right)^{2/3}\times\nonumber\\
\times\int_0^\tau\frac{\mathrm{d}\tau_1}{\sinh^2\tau_1}
\int_0^{\tau_1}\kappa_1^2(\tau_2)\,\sinh^{2/3}\tau_2\,\mathrm{d}\tau_2
\end{eqnarray}
where once again $\delta a_0(t\rightarrow 0)=0$.
Lastly, we compute the correction to the scale factor along the $z$ direction, $\delta b_0$, from \eqref{equdeltab0}:
\begin{eqnarray}
\delta b_0(t)=-2\,\delta a_0(t)+\frac{\bar\kappa_1^2}{10\,\Omega_\Lambda}\left(\frac{\Omega_\Lambda}{1-\Omega_\Lambda}\right)^{2/3}\times\nonumber\\
\times\phantom{.}_2F_1\left(\frac{3}{2},\frac{5}{6},\frac{11}{6},-\sinh^2\tau\right)\cosh\tau\sinh^{2/3}\tau\,\,.
\end{eqnarray}
\section{The perturbed redshift}\label{rs
An observable quantity related to the metric found in the previous section is the angular dependence of the CMB radiation induced by the integrated Sachs-Wolfe effect on our system. We compute it in this and in the following section. Photons coming from the last scattering surface move along null geodesics:
\begin{equation}
\frac{\mathrm{d} k^\mu}{\mathrm{d}\lambda}+\Gamma^\mu_{\nu\rho}\,k^\nu\,k^\rho=0\,\,,
\end{equation}
where $\lambda$ is the affine parameter along the worldline $x^\mu(\lambda)$ of the photon. $k^\mu=\mathrm{d} x^\mu/\mathrm{d}\lambda$ denotes the tangent (null) vector to the photon worldline.
It is convenient to introduce the conformal time $\eta$, so that the metric reads
\begin{eqnarray}
\mathrm{d} s^2=&&a_0^2(\eta)\,\left[-\mathrm{d}\eta^2+(1+2\,\delta a(\eta,\,z))\,\left(\mathrm{d} x^2+\mathrm{d} y^2\right)+\right.\nonumber\\
&&\left.+(1+2\,\delta b(\eta,\,z))\,\mathrm{d} z^2\right]\,\,.
\end{eqnarray}
This metric will allow us to factor out the dependence of the redshift $\zeta$~\footnote{We denote the redshift by $\zeta$ because $z$ is already chosen for one of the coordinates.} on the scale factor so to effectively simplify the calculation. Let us perform the conformal transformation
\begin{equation}\label{conformal}
\mathrm{d}{s}^2=a_0^2(\eta)\,\mathrm{d}\bar{s}^2\,,\ \mathrm{d}{\lambda}=a_0^2(\eta)\,\mathrm{d}{\bar\lambda}\,,\ {k}^\mu=a_0(\eta)^{-2}\bar{k}^\mu\,\,.
\end{equation}
It is easy to show by taking advantage of $k_\mu$ being a null vector, that $\bar k^\mu$ satisfies
\begin{equation}
\frac{\mathrm{d}\bar k^\mu}{\mathrm{d}\bar\lambda}+\bar\Gamma^\mu_{\nu\rho}\,\bar k^\nu\,\bar k^\rho=0\,\,,
\end{equation}
where $\bar\Gamma^\mu_{\nu\rho}$ are the Christoffels symbols of the metric $\mathrm{d}\bar s^2$, which is a first order perturbation around the Minkowski metric.
The photon worldline, $x^\alpha(\bar\lambda)$, can be expanded in a perturbation series
\begin{equation}
x^\mu(\bar\lambda)=x_B^\mu(\bar\lambda)+\delta x^\mu(\bar\lambda)\,,\ \bar{k}^\mu(\bar\lambda)=\bar{k}_B^\mu(\bar\lambda)+\delta\bar{k}^\mu(\bar\lambda)\,\,,
\end{equation}
where the subscript $B$ denotes a background quantity and $\delta$ denotes the first-order perturbation on it.
Since the background component of $\mathrm{d}\bar s^2$ describes a Minkowski spacetime, then the background null vector $\bar{k}_B^\mu$ is constant. By appropriately scaling $\bar\lambda$ we impose $\bar k_B^0=-1$. We have $\bar{k}_B^\mu=x_B^\mu{}'(\bar\lambda)=(-1,\,\vec{n})$, where $\vec{n}$ is a unit 3-vector, as we will eventually choose $\bar\lambda=0$ for the observer of the CMB photon, with $\lambda$ increasing as we go back in time. Given the symmetries of our spacetime, we can locate the observer at $(x_B)_O=(y_B)_O=0$, while $(z_B)_O=z_0$ remains arbitrary, so we have
\begin{eqnarray}\label{flatsol}
&&\eta_B(\bar\lambda)=\eta_0-\bar\lambda\,\,,\\
&&x_B(\bar\lambda)=n_x\,\bar\lambda\,,\ y_B(\bar\lambda)=n_y\,\bar\lambda\,,\ z_B(\bar\lambda)=z_0+n_z\,\bar\lambda\,\,.\nonumber
\end{eqnarray}
The redshift of a source with four-velocity $u^\mu_S$, as seen by an observer whose four-velocity is $u^\mu_O$, is defined as
\begin{equation}\label{redshift}
1+\zeta=\frac{{({g}_{\mu\nu} {k}^\mu {u}^\nu)}_{S}}
{{({g}_{\alpha\beta} {k}^\alpha {u}^\beta)}_{O}}\,\,.
\end{equation}
It is convenient to introduce a ``conformal'' redshift $\bar\zeta$, defined by adding a bar to all the quantities in eq.~\eqref{redshift} and defining $u^\mu=a_0^{-1}\,(\eta) \bar{u}^\mu$:
\begin{equation}\label{confredsh}
1+\zeta=\frac{1+\bar\zeta} {a_0(\eta_S)}\,\,,
\end{equation}
where we have used the fact that $a_0(\eta_0)=1$. We assume that the spatial components of $\bar u^\mu_S$ and $\bar u^\mu_O$ can be treated as first order quantities, in which case $\bar\zeta$ is also a first order quantity. By using $\bar{u}^\mu_{O,\,S}=(1,\,\vec{v}_{O,\,S})$, where $\vec{v}_O$ is the peculiar velocity of the observer and ${\vec{v}}_S$ of the source, we compute $\bar\zeta$ to be
\begin{equation}
\bar\zeta\approx\left({\vec{v}}_S-{\vec{v}}_O\right)\cdot\vec{n}- \delta \bar{k}^0(\bar\lambda_S)+\delta \bar{k}^0(\bar\lambda=0)\,\,,
\end{equation}
where we have used the fact that the photon was emitted at $\bar\lambda=\bar\lambda_S$ and is observed at $\bar\lambda=0$. The quantity $\delta \bar{k}^0({\bar{\lambda}_S})-\delta \bar{k}^0(0)$ is obtained by integrating the geodesic equation for $\bar k^0$
\begin{equation}
\frac{\mathrm{d}\,\delta\bar k^0}{\mathrm{d}\bar\lambda}+\frac{\partial\,\delta a}{\partial\eta}\left(n_x^2+n_y^2\right)+\frac{\partial\,\delta b}{\partial\eta}\,n_z^2=0\,\,.
\end{equation}
Therefore $\bar\zeta$ is
\begin{equation}\label{finalz}
\bar\zeta\approx\left(\vec{v}_S-\vec{v}_O\right)\cdot\vec{n}+
\int_0^{\bar{\lambda}_S}\left[\frac{\partial\,\delta a}{\partial\eta}\left(n_x^2+n_y^2\right)+\frac{\partial\,\delta b}{\partial\eta}\,n_z^2\right]\,\mathrm{d}\bar\lambda\,\,,
\end{equation}
where the first term corresponds to the intrinsic motions of the source and of the observer. The second term is associated to the integrated Sachs-Wolfe effect on our spacetime; the next section will be dedicated to a thorough discussion of this term.
\section{CMB anisotropies}\label{cmb
The effect of our anisotropic space on the CMB spectrum is given by the integrated Sachs-Wolfe (iSW) effect:
\begin{equation}\label{zeta}
\bar\zeta_{\mathrm{iSW}}\approx \int_0^{\bar\lambda_S}\left[\frac{\partial\,\delta {a}}{\partial\eta}\,(n_x^2+n_y^2)+\frac{\partial\,\delta {b}}{\partial\eta}\,n_z^2\right]\,\mathrm{d}\bar{\lambda}\,\,.
\end{equation}
Using~\eqref{flatsol}, we trade $\bar\lambda$ for $\eta$, moreover, since $\delta a$ and $\delta b$ rapidly vanish as $\eta\rightarrow 0$, we can extend the integration to $\eta=0$ ({\em i.e.}, $\bar\lambda_S=\eta_0$). We remind that $\delta a\approx\delta a_0(\eta)+\delta a_2(\eta)\,z^2$ and $\delta b\approx\delta b_0(\eta)$. Since $\delta a_2$ and $\delta b_2$ appear on the same footing in the equations above, and since $\delta b_2$ is undetermined at this level of approximation, we neglect also $\delta a_2$ in what follows. Thus we have
\begin{equation}
\bar\zeta_{\mathrm{iSW}}\approx \int_0^{\eta_0}\left(\frac{\mathrm{d}\,\delta a_0}{\mathrm{d}\eta}\sin^2\vartheta+\frac{\mathrm{d}\,\delta b_0}{\mathrm{d}\eta}\cos^2\vartheta\right)\,\mathrm{d}\eta\,\,,
\end{equation}
where $n_z\equiv\cos\vartheta$.
The above expression can be integrated explicitly
\begin{equation}
\bar\zeta_{\mathrm{iSW}}\approx \delta a_0(\eta_0)+\left(\delta b_0(\eta_0)-\delta a_0(\eta_0)\right)\,\cos^2\vartheta\,\,,
\end{equation}
in which we have used $\delta a_0(\eta=0)=\delta b_0(\eta=0)=0$. It is assumed that the direction of alignment of the gradient coincides precisely with the $z$ axis. In order to find the equivalent expression when the gradient of $\phi$ is directed along a direction $(\Theta,\,\Phi)$, we rotate the coordinate system $(x,\,y,\,z)$ by an angle $\Phi$ around the $z$ axis and by an angle $\Theta$ around the $x$ axis. We thus obtain
\begin{eqnarray}
\bar\zeta_{\mathrm{iSW}}(\vartheta,\,\varphi)&&\approx \delta a_0(\eta_0)+\left(\delta b_0(\eta_0)-\delta a_0(\eta_0)\right)\times\\
&&\times\left[\sin\vartheta\,\sin\Theta\,\cos(\Phi+\varphi)+\cos\vartheta\,\cos\Theta\right]^2\,\,.\nonumber
\end{eqnarray}
We follow, {\em e.g.}, \cite{Koivisto:2008ig} to find the effect on CMB. If $T_*$ is the CMB temperature at decoupling, then the observed temperature will be $T(\vartheta,\,\varphi)=T_*/(1+\zeta(\vartheta,\,\varphi))\simeq T_*\,a_{\mathrm{CMB}}\,(1-\bar\zeta(\vartheta,\,\varphi))$, where $a_{\mathrm{CMB}}$ is the scale factor at decoupling. Then the average observed temperature is $\langle T\rangle=\int\mathrm{d}\varphi\,\mathrm{d}\cos\vartheta\,T(\vartheta,\,\varphi)/4\pi$ and the anisotropy $\delta T(\vartheta,\,\varphi)/T=1-T(\vartheta,\,\varphi)/\bar{T}=\bar\zeta_{\mathrm{iSW}}(\vartheta,\,\varphi)-\int\mathrm{d}\varphi\,\mathrm{d}\cos\vartheta\,\bar\zeta_{\mathrm{iSW}}(\vartheta,\,\varphi)/4\pi$. Its decomposition in spherical harmonics is given by
\begin{equation}
a_{\ell m}=\int\mathrm{d}\varphi\,\mathrm{d}\cos\vartheta\frac{\delta T(\vartheta,\,\varphi)}{T}\,Y_\ell^m{}^*\,\,.
\end{equation}
An explicit calculation allows to show that the only non-vanishing contributions from the anisotropic integrated Sachs-Wolfe effect are
\begin{eqnarray}
a^{\mathrm{iSW}}_{22}&=&\sqrt{\frac{2\,\pi}{15}}\,\left(\delta b_0(\eta_0)-\delta a_0(\eta_0)\right)\,\sin^2\Theta\,{\mathrm{e}}^{2i\Phi}\,\,,\\
a^{\mathrm{iSW}}_{21}&=&-\sqrt{\frac{2\,\pi}{15}}\,\left(\delta b_0(\eta_0)-\delta a_0(\eta_0)\right)\,\sin\,2\Theta\,{\mathrm{e}}^{i\Phi}\,\,,\nonumber\\
a^{\mathrm{iSW}}_{20}&=&\frac{1}{3}\,\sqrt{\frac{\pi}{5}}\,\left(\delta b_0(\eta_0)-\delta a_0(\eta_0)\right)\,\left(1+3\,\cos\,2\Theta\right)\,\,,\nonumber
\end{eqnarray}
with $a_{2,-1}=-a_{21}^*$ and $a_{2,-2}=a_{22}^*$.
These contributions have to be summed to the intrinsic quadrupole momentum of the CMB. It is possible to see that our contribution can help explaining the low quadrupole observed in CMB data, since it effectively gives rise to an ``ellipsoidal'' Universe~\cite{Campanelli:2007qn}.
Let us quickly review the argument of~\cite{Campanelli:2007qn}: we denote by $a_{2m}^I$ (with $a^I_{2,-m}=(-1)^m\,a^I_{2m}{}^*$) the quadrupole components of the ``intrinsic'' CMB fluctuations. As a consequence of statistical isotropy, the $a^I_{2m}$ can be assumed to be equal up to a phase, {\em i.e.}, $a^I_{20}=\sqrt{\pi/3}\,{\cal Q}_I$, $a^I_{21}=\sqrt{\pi/3}\,{\mathrm{e}}^{i\,\alpha_1}\,{\cal Q}_I$, $a^I_{22}=\sqrt{\pi/3}\,{\mathrm{e}}^{i\,\alpha_2}\,{\cal Q}_I$.
The quadrupole amplitude is defined by $Q_2^2=\frac{3}{5\,\pi}\sum_m\left|a^{\mathrm{iSW}}_{2m}+a_{2m}^I\right|^2$. If $\Theta$, $\Phi$ and $\left(\delta b_0(\eta_0)-\delta a_0(\eta_0)\right)$ are appropriately chosen, then one can obtain a ``low'' value of $Q_2\simeq 5.3\times 10^{-6}$ even if the prediction for a spherical Universe, ${\cal {Q}}_I\simeq 1.3\times 10^{-5}$, would be higher. In particular, from the analysis of~\cite{Campanelli:2006vb}, we can extract the required value of $\left|\delta b_0(\eta_0)-\delta a_0(\eta_0)\right|\simeq 2\times 10^{-5}$.
It is however important to note that the cancellation of the intrinsic quadrupole component by the integrated Sachs-Wolfe effect associated to the ellipsoidal Universe requires a correlation between the primordial and the late components, as discussed already in~\cite{Contaldi:2003zv} and quantified in~\cite{Gruppuso:2007ya,Appleby:2009za}. Even in the absence of correlations with the intrinsic quadrupole, the effect on the CMB provides the strongest constraints on our scenario: those constraints can be expressed as $\left|\delta b_0(\eta_0)-\delta a_0(\eta_0)\right|\lesssim 2\times 10^{-5}$. In figure~\ref{fig1}, we show the log-plot of $\left|\delta b_0(\eta_0)-\delta a_0(\eta_0)\right|/\bar\kappa_1^2$ as a function of the parameter $\mu$. This plot shows how, for $m\gtrsim {\cal {O}}(\mathrm{few})\times H_0$, the degree of anisotropy grows exponentially with $m$. Note also that, even for $m=0$, we can have a small anisotropy imprinted on the CMB. As we will discuss below, however, in this case the original gradient in $\phi$ would not have been negligible during inflation, and its effect should have been taken into account in the calculation of the primordial spectrum of density fluctuations.
\begin{figure}[]
\begin{center}
\includegraphics[width=3in]{log_b0a0_vs_mu.pdf}
\caption{The quantity $|\delta b_0(\eta_0)-\delta a_0(\eta_0)|/\bar\kappa_1^2$ plotted as a function of $\mu$.}
\label{fig1}
\end{center}
\end{figure}
\section{On the initial gradient}\label{gradient
In this section we discuss the magnitude of the initial gradient $\kappa_1(0)\equiv \bar\kappa_1\,H_0\,M_\mathrm{Pl}$. Since for $H\gg m$ the evolution of the field $\phi$ is frozen by Hubble friction, we have $\phi\simeq \kappa_1(0)\,z$ inside our Hubble patch. As a consequence, the energy $\rho_\phi=T_{tt}^\phi$ scales as $\kappa_1(0)^2/\left(2\,a_0(t)^2\right)$. This gradient term redshifts as $a(t)^{-2}$, {\em i.e.}, as spatial curvature. During inflation, with Hubble parameter $H_I$, $\phi$ provides a fraction
\begin{equation}
\frac{\rho_\phi}{\rho_{\mathrm{tot}}}\simeq \frac{\bar\kappa_1^2}{6}\,\frac{H_0^2}{H_I^2}\,\frac{{\mathrm{e}}^{2\,N}}{a_{\mathrm{end}}^2}
\end{equation}
of the background inflaton energy, where $N$ is the number of e-foldings from the end of inflation and $a_{\mathrm{end}}$ is the scale factor at the end of inflation. In particular, if inflation lasted $N_{\mathrm{i}}$ e-foldings and we denote by $T_{RH}\simeq (3\,M_\mathrm{Pl}^2\,H_I^2)^{1/4}$ the reheating temperature (we assume instantaneous reheating and ignore the effects of $g_*$, the effective number of relativistic degrees of freedom in the system), then we can trade $\kappa_1(0)$ for $(\rho_\phi/\rho_{\mathrm{tot}})|_\mathrm{i}$ computed at the beginning of inflation
\begin{equation}
\bar\kappa_1\simeq \sqrt{2\left.\frac{\rho_\phi}{\rho_\mathrm{tot}}\right|_\mathrm{i}}\,\frac{T_0\,T_{RH}}{H_0\,M_\mathrm{Pl}}\,{\mathrm{e}}^{-N_{\mathrm{i}}}\,\,,
\end{equation}
where we have used $a_{\mathrm{end}}=T_0/T_{RH}$ with $T_0\simeq 2\times 10^{-4}$eV being the current value of the CMB temperature.
As discussed in section~\ref{cmb}, data require $|\delta b_0(\eta_0-\delta a_0(\eta_0)|\lesssim 2\times 10^{-5}$, that implies
\begin{equation}
\left.\frac{\rho_\phi}{\rho_\mathrm{tot}}\right|_\mathrm{i}\lesssim 10^{-5}\,\left[\frac{\left|\delta b_0(\eta_0)-\delta a_0(\eta_0)\right|}{\bar\kappa_1^2}\right]^{-1}\,\frac{H_0^2\,M_\mathrm{Pl}^2}{T_0^2\,T_{RH}^2}\,{\mathrm{e}}^{2\,N_\mathrm{i}}\,\,,
\end{equation}
where the quantity $\left|\delta b_0(\eta_0)-\delta a_0(\eta_0)\right|/\bar\kappa_1^2$ is plotted in figure~\ref{fig1}.
Finally, we use the relation ${\mathrm{e}}^{N_{\mathrm{obs}}}\simeq (H_I\,T_0)/(H_0\,T_{RH})$ ($N_{\mathrm{obs}}$ is the number of observable e-foldings of inflation) to obtain the condition
\begin{equation}
\left.\frac{\rho_\phi}{\rho_\mathrm{tot}}\right|_\mathrm{i}\simeq 3\times 10^{-6}\,\left[\frac{\left|\delta b_0(\eta_0)-\delta a_0(\eta_0)\right|}{\bar\kappa_1^2}\right]^{-1}\,{\mathrm{e}}^{2\,(N_{\mathrm{i}}-N_{\mathrm{obs}})}\,\,.
\end{equation}
Therefore inflation can start with a sizable (but still smaller than unity) value of $(\rho_\phi/\rho_\mathrm{tot})|_\mathrm{i}$ provided $N_\mathrm{i}$ is sufficiently larger than $N_{\mathrm{obs}}$. Then when the observable scales left the horizon, $N_{\mathrm{obs}}$ e-foldings before the end of inflation, the perturbation in the scalar $\phi$ was a smaller fraction (by a factor $\exp\left[2\,(N_{\mathrm{obs}}-N_i)\right]$) of the background energy, small enough to affect neither the dynamics of the inflaton nor that of the perturbations. The gradient in $\phi$ stays irrelevant until the Hubble parameter is comparable with $m$, when it will start increasing under the effect of the tachyonic potential and will give a perturbation of the right amplitude today.
To fix ideas we can set $N_{\mathrm{obs}}=60$, $\mu=10$ (so that $\left|\delta b_0(\eta_0)-\delta a_0(\eta_0)\right|/\bar\kappa_1^2\simeq 10^5$), $N_\mathrm{i}=71$. Then, $(\rho_\phi/\rho_\mathrm{tot})|_\mathrm{i}\simeq 0.1$. At the time observable scales exited the horizon, $\rho_\phi/\rho_{\mathrm{tot}}$ would have redshifted by a factor ${\mathrm{e}}^{2\times (60-71)}\simeq 10^{-10}$ and it would have been negligible. In the absence of the tachyonic mass, the gradient of $\phi$ would be still negligible today. However, due to the tachyonic enhancement $\left|\delta b_0(\eta_0)-\delta a_0(\eta_0)\right|$, the effects of such a primordial gradient will manifest themselves in the lowest CMB multipoles~\footnote{The possibility of a gradient in the dark energy is also considered in~\cite{Gordon:2005ai}, where, however, it is assumed that the quintessence is so light that its dynamics is frozen.}.
Before concluding this section, let us discuss the possible worry that, since $\phi$ is effectively a massless field for most of the history of the Universe, its quantum fluctuations generated during inflation may be larger than the classical gradient that is central to our analysis. It is easy to choose the inflationary parameters in such a way that the quantum fluctuations are subdominant with respect to the original gradient. Indeed, the amplitude of the quantum fluctuations in $\phi$ will be of the order of $H_I$, the Hubble parameter during inflation. During inflation, the modulation of $\phi$ at scales $z\simeq H_0^{-1}$, which are relevant today, is of the order of $\kappa_1(0)\,z\simeq \bar\kappa_1\,M_\mathrm{Pl}\simeq \sqrt{(\rho_\phi/\rho_\mathrm{tot})|_\mathrm{i}}\,{\mathrm{e}}^{N_{\mathrm{obs}}-N_\mathrm{i}}\,M_\mathrm{Pl}$. Assuming $(\rho_\phi/\rho_\mathrm{tot})|_\mathrm{i}={\cal {O}}(1)$, we see that, as long as $H_I\ll M_\mathrm{Pl}\,{\mathrm{e}}^{N_{\mathrm{obs}}-N_\mathrm{i}}$ ({\em e.g.}, $H_I\ll 10^{-5}\,M_\mathrm{Pl}$ in the example $N_\mathrm{i}-N_{\mathrm{obs}}=11$ considered above), the effects of quantum fluctuations are negligible.
\section{Conclusions and Discussion
If inflation lasted a relatively short time, then some primordial gradients could exist as a relic of the chaotic pre-inflationary dynamics. A gradient in the quintessence field, even if too small to leave any detectable effects during the observable epoch of inflation, might be amplified by a tachyonic quintessence to have observable magnitude today. We have seen that in the simplest scenario, the dominant effect of the amplified gradient is on the CMB quadrupole, whose observed amplitude sets the strongest constraints on the parameters of the problem.
The fact that CMB fluctuations are affected only at the quadrupole level is a consequence of the approximation $\phi\simeq \kappa_1(t)\,z$. Higher powers of $z$ in the expansion of $\phi$ would lead to the generation of higher multipole contribution. Generically, terms of order $z^n$ in the expansion of the scalar field $\phi(t,z)$ will generate contributions on both $\delta a(t,z)$ and $\delta b(t,z)$ up to order $z^{n-1}$, which, in turn, provide terms up to order $z^{n+1}$ in $\bar\zeta_\mathrm{iSW}$ as for equation~\eqref{zeta}. That is, terms of order $z^n$ in $\phi(t,z)$ will affect $C_\ell$'s of $\ell=n+1$. Given the hints of multipole alignments up to $\ell\sim 40$, it would be interesting to be able to extend our mechanism in such a way that larger powers of $z$ in the expression of $\phi(t,\,z)$ are generated. One possibility is that the self-interactions of $\phi$ during the recent cosmological evolution are responsible. For instance, if the quintessence field $\phi$ is given by a pseudo-Nambu-Goldstone boson, its potential is $V(\phi)=\Lambda\,\left[1+\cos\left(\phi/f\right)\right]/2$ ($f^2=\Lambda/(2\,m^2)$). By following our analysis of section~\ref{metric}, we see that, because of the Klein-Gordon equation, the coefficients of lower powers in $z$ will act as sources for those of higher powers, hence generating higher multipoles. A further investigation of this mechanism is subject of future work.\\
{\bf Acknowledgments.} This work has been supported in part by the NSF grant PHY - 0855119. We thank John Donoghue, Burak Himmetoglu, David Langlois, Nemanja Kaloper and Marco Peloso for interesting discussions.
\bibliographystyle{apsrev}
|
\section{Introduction}
\label{sec:intro}
\subsection{Concise signal models}
A significant byproduct of the Information Age has been an explosion
in the sheer quantity of raw data demanded from sensing systems.
From digital cameras to mobile devices, scientific computing to
medical imaging, and remote surveillance to signals intelligence,
the size (or dimension) $\dim$ of a typical desired signal continues
to increase. Naturally, the dimension $\dim$ imposes a direct burden
on the various stages of the data processing pipeline, from the data
acquisition itself to the subsequent transmission, storage, and/or
analysis; and despite rapid and continual improvements in computer
processing power, other bottlenecks do remain, such as communication
bandwidth over wireless channels, battery power in remote sensors
and handheld devices, and the resolution/bandwidth of
analog-to-digital converters.
Fortunately, in many cases, the information contained within a
high-dimensional signal actually obeys some sort of concise,
low-dimensional model. Such a signal may be described as having just
$\sparsity \ll \dim$ degrees of freedom for some $\sparsity$.
Periodic signals bandlimited to a certain frequency are one example;
they live along a fixed $\sparsity$-dimensional linear subspace of
$ \mathbb{R} ^\dim$. Piecewise smooth signals are an example of {\em sparse
signals}, which can be written as a succinct linear combination of
just $\sparsity$ elements from some basis such as a wavelet
dictionary. Still other signals may live along $\mdim$-dimensional
submanifolds of the ambient signal space $ \mathbb{R} ^\dim$; examples
include collections of signals observed from multiple viewpoints in
a camera or sensor network. In general, the conciseness of these
models suggests the possibility for efficient processing and
compression of these signals.
\subsection{Compressive measurements}
Recently, the conciseness of certain signal models has led to the
use of {\em compressive measurements} for simplifying the data
acquisition process. Rather than designing a sensor to measure a
signal $x \in \mathbb{R} ^\dim$, for example, it often suffices to design
a sensor that can measure a much shorter vector $y = \proj x$, where
$\proj$ is a linear measurement operator represented as an $\pdim
\times \dim$ matrix, and where typically $\pdim \ll \dim$. As we
discuss below in the context of Compressive Sensing (CS), when
$\proj$ is properly designed, the requisite number of measurements
$\pdim$ typically scales with the information level $\sparsity$ of
the signal, rather than with its ambient dimension $\dim$.
Surprisingly, the requirements on the measurement matrix $\proj$ can
often be met by choosing $\proj$ randomly from an acceptable
distribution. One distribution allows the entries of $\proj$ to be
chosen as i.i.d.\ Gaussian random variables; another dictates that
$\proj$ has orthogonal rows that span a random $\pdim$-dimensional
subspace of $ \mathbb{R} ^\dim$.
Physical architectures have been proposed for hardware that will
enable the acquisition of signals using compressive
measurements~\cite{duarte2008spi,candes2008ics,healy2008cpi,demod}.
The potential benefits for data acquisition are numerous. These
systems can enable simple, low-cost acquisition of a signal directly
in compressed form without requiring knowledge of the signal
structure in advance. Some of the many possible applications include
distributed source coding in sensor networks~\cite{dcsJournal},
medical imaging~\cite{lustig2008csm}, high-rate analog-to-digital
conversion~\cite{candes2008ics,healy2008cpi,demod}, and error
control coding~\cite{CandesDLP}.
\subsection{Signal understanding from compressive measurements}
Having acquired a signal $x$ in compressed form (in the form of a
measurement vector $y$), there are many questions that may then be
asked of the signal. These include:
\begin{itemize}
\item [Q1.] {\em Recovery:} What was the original signal $x$?
\item [Q2.] {\em Sketching:} Supposing that $x$ was sparse or nearly so,
what were the $\sparsity$ basis vectors used to generate $x$?
\item [Q3.] {\em Parameter estimation:} Supposing $x$ was generated from a
$\mdim$-dimensional parametric model, what was the original
$\mdim$-dimensional parameter that generated $x$?
\end{itemize}
Given only the measurements $y$ (possibly corrupted by noise),
solving any of the above problems requires exploiting the concise,
$\sparsity$-dimensional structure inherent in the
signal.\footnote{Other problems, such as finding the nearest
neighbor to $x$ in a large database of signals~\cite{JL_Indyk}, can
also be solved using compressive measurements and do not require
assumptions about the concise structure in $x$.} CS addresses
questions Q1 and Q2 under the assumption that the signal $x$ is
$\sparsity$-sparse (or approximately so) in some basis or
dictionary; in Section~\ref{sec:cs} we outline several key
theoretical bounds from CS regarding the accuracy to which these
questions may be answered.
\subsection{Manifold models for signal understanding}
\label{sec:mmodelsunder}
In this paper, we will address these questions in the context of a
different modeling framework for concise signal structure. Instead
of sparse models, we focus on the broad class of {\em manifold
models}, which arise both in settings where a $\mdim$-dimensional
parameter $\theta$ controls the generation of the signal and also in
non-parametric settings.
As a very simple illustration, consider the articulated signal in
Figure~\ref{fig:CSM}(a). We let $g(t)$ be a fixed continuous-time
Gaussian pulse centered at $t=0$ and consider a shifted version of
$g$ denoted as the parametric signal $f_\theta(t) := g(t-\theta)$
with $t,\theta \in [0,1]$. We then suppose the discrete-time signal
$x = x_\theta \in \mathbb{R} ^\dim$ arises by sampling the continuous-time
signal $f_\theta(t)$ uniformly in time, i.e., $x_\theta(n) =
f_\theta(n/\dim)$ for $n=1,2,\dots,\dim$. As the parameter $\theta$
changes, the signals $x_\theta$ trace out a continuous
one-dimensional (1-D) curve $\manifold = \{x_\theta: \theta \in
[0,1]\} \subset \mathbb{R} ^\dim$. The conciseness of our model (in
contrast with the potentially high dimension $\dim$ of the signal
space) is reflected in the low dimension of the path $\manifold$.
\begin{figure}[t]
\begin{center}
\epsfysize = 48mm \epsffile{fGauss.eps} \quad\quad
\epsfysize = 48mm \epsffile{gaussianEmbed.eps} \quad\quad\quad
\end{center}
\vspace*{-5mm} \caption{\small\sl \label{fig:CSM} (a) The
articulated signal $f_\theta(t) = g(t-\theta)$ is defined via shifts
of a primitive function $g$, where $g$ is a Gaussian pulse. Each
signal is sampled at $\dim$ points, and as $\theta$ changes, the
resulting signals trace out a 1-D manifold in $ \mathbb{R} ^\dim$. (b)
Projection of the manifold from $ \mathbb{R} ^\dim$ onto a random 3-D
subspace; the color/shading represents different values of $\theta
\in [0,1]$.}
\end{figure}
In the real world, manifold models may arise in a variety of
settings. A $\mdim$-dimensional parameter $\theta$ could reflect
uncertainty about the 1-D timing of the arrival of a signal (as in
Figure~\ref{fig:CSM}(a)), the 2-D orientation and position of an
edge in an image, the 2-D translation of an image under study, the
multiple degrees of freedom in positioning a camera or sensor to
measure a scene, the physical degrees of freedom in an articulated
robotic or sensing system, or combinations of the above. Manifolds
have also been proposed as approximate models for signal databases
such as collections of images of human faces or of handwritten
digits~\cite{Eigenfaces,digits,broomhead01wh}.
Consequently, the potential applications of manifold models are
numerous in signal processing. In some applications, the signal $x$
itself may be the object of interest, and the concise manifold model
may facilitate the acquisition or compression of that signal.
Alternatively, in parametric settings one may be interested in using
a signal $x = x_\theta$ to infer the parameter $\theta$ that
generated that signal.
In an application known as manifold learning, one may be presented
with a collection of data $\{x_{\theta_1}, x_{\theta_2}, \dots,
x_{\theta_n}\}$ sampled from a parametric manifold and wish to
discover the underlying parameterization that generated that
manifold.
Multiple manifolds can also be considered simultaneously, for
example in problems that require recognizing an object from one of
$n$ possible classes, where the viewpoint of the object is uncertain
during the image capture process. In this case, we may wish to know
which of $n$ manifolds is closest to the observed image $x$.
While any of these questions may be answered with full knowledge of
the high-dimensional signal $x \in \mathbb{R} ^\dim$, there is growing
theoretical and experimental support that they can also be answered
from only compressive measurements $y = \proj x$. In a recent paper,
we have shown that given a sufficient number $\pdim$ of random
measurements, one can ensure with high probability that a manifold
$\manifold \subset \mathbb{R} ^\dim$ has a stable embedding in the
measurement space $ \mathbb{R} ^\pdim$ under the operator $\proj$, such
that pairwise Euclidean and geodesic distances are approximately
preserved on its image $\proj \manifold$. We restate the precise
result in Section~\ref{sec:manifolds}, but a key aspect is that the
number of requisite measurements $\pdim$ is linearly proportional to
the information level of the signal, i.e., the dimension $\mdim$ of
the manifold.
As a very simple illustration of this embedding phenomenon,
Figure~\ref{fig:CSM}(b) presents an experiment where just $\pdim =
3$ compressive measurements are acquired from each point $x_\theta$
described in Figure~\ref{fig:CSM}(a). We let $\dim = 1024$ and
construct a randomly generated $3 \times \dim$ matrix $\proj$ with
orthogonal rows. Each point $x_\theta$ from the original manifold
$\manifold \subset \mathbb{R} ^{1024}$ maps to a unique point $\proj
x_\theta$ in $ \mathbb{R} ^3$; the manifold embeds in the low-dimensional
measurement space. Given any $y = \proj x_{\theta'}$ for $\theta'$
unknown, then, it is possible to infer the value $\theta'$ using
only knowledge of the parametric model for $\manifold$ and the
measurement operator $\proj$. Moreover, as the number $\pdim$ of
compressive measurements increases, the manifold embedding becomes
much more stable and remains highly self-avoiding.
Indeed, there is strong empirical evidence that, as a consequence of
this phenomenon, questions such as Q1 (signal recovery) and Q3
(parameter estimation) can be accurately solved using only
compressive measurements of a signal $x$, and that these procedures
are robust to noise and to deviations of the signal $x$ away from
the manifold $\manifold$~\cite{mbwPhdThesis,davenport2007sfc}.
Additional theoretical and empirical justification has followed for
the manifold learning~\cite{hegde2007rpm} and multiclass recognition
problems~\cite{davenport2007sfc} described above.
Consequently, many of the advantages of compressive measurements
that are beneficial in sparsity-based CS (low-cost sensor design,
reduced transmission requirements, reduced storage requirements,
lack of need for advance knowledge of signal structure, simplified
computation in the low-dimensional space $ \mathbb{R} ^\pdim$, etc.)\ may
also be enjoyed in settings where manifold models capture the
concise signal structure. Moreover, the use of a manifold model can
often capture the structure of a signal in many fewer degrees of
freedom $\mdim$ than would be required in any sparse representation,
and thus the measurement rate $\pdim$ can be greatly reduced
compared to sparsity-based CS approaches.
In this paper, we will focus on questions Q1 (signal recovery) and
Q3 (parameter estimation) and reinforce the existing empirical work
by establishing theoretical bounds on the accuracy to which these
questions may be answered.
We will consider both deterministic and probabilistic
instance-optimal bounds, and we will see strong similarities to
analogous results that have been derived for sparsity-based CS. As
with sparsity-based CS, we show for manifold-based CS that for any
fixed $\proj$, uniform deterministic $\ell_2$ recovery bounds for
recovery of all $x$ are necessarily poor. We then show that, as with
sparsity-based CS, providing for any $x$ a probabilistic bound that
holds over most $\proj$ is possible with the desired accuracy. We
consider both noise-free and noisy measurement settings and compare
our bounds with sparsity-based CS.
\subsection{Paper organization}
We begin in Section~\ref{sec:cs} with a brief review of CS topics,
to set notation and to outline several key results for later
comparison. In Section~\ref{sec:manifolds} we discuss manifold
models in more depth, restate our previous bound regarding stable
embeddings of manifolds, and formalize our criteria for answering
questions Q1 and Q3 in the context of manifold models. In
Section~\ref{sec:deter}, we confront the task of deriving
deterministic instance-optimal bounds in $\ell_2$. In
Section~\ref{sec:prob}, we consider instead probabilistic
instance-optimal bounds in $\ell_2$. We conclude in
Section~\ref{sec:concl} with a final discussion.
\section{Sparsity-Based Compressive Sensing} \label{sec:cs}
\subsection{Sparse models}
The concise modeling framework used in Compressive Sensing (CS) is
{\em sparsity}. Consider a signal $x \in \mathbb{R} ^\dim$ and suppose the
$\dim \times \dim$ matrix $\Psi = [\psi_1 ~ \psi_2 ~ \cdots~
\psi_N]$ forms an orthonormal basis for $ \mathbb{R} ^\dim$. We say $x$ is
$\sparsity$-sparse in the basis $\Psi$ if for $\alpha \in
\mathbb{R} ^\dim$ we can write
$$
x = \Psi \alpha,
$$
where $\|\alpha\|_0 = \sparsity < \dim$. (The $\ell_0$-norm notation
counts the number of nonzeros of the entries of $\alpha$.) In a
sparse representation, the actual information content of a signal is
contained exclusively in the $\sparsity < \dim$ positions and values
of its nonzero coefficients.
For those signals that are approximately sparse, we may measure
their proximity to sparse signals as follows. We define
$\alpha_\sparsity \in \mathbb{R} ^\dim$ to be the vector containing only
the largest $\sparsity$ entries of $\alpha$, with the remaining
entries set to zero. Similarly, we let $x_\sparsity = \Psi
\alpha_\sparsity$. It is then common to measure the proximity to
sparseness using either $\|\alpha-\alpha_\sparsity\|_1$ or
$\norm{\alpha-\alpha_\sparsity}$ (the latter of which equals
$\norm{x-x_\sparsity}$ because $\Psi$ is orthonormal).
\subsection{Compressive measurements}
\label{sec:csmeas}
CS uses the concept of sparsity to simplify the data acquisition
process. Rather than designing a sensor to measure a signal $x \in
\mathbb{R} ^\dim$, for example, it often suffices to design a sensor that
can measure a much shorter vector $y = \proj x$, where $\proj$ is a
linear measurement operator represented as an $\pdim \times \dim$
matrix, and typically $\pdim \ll \dim$.
The measurement matrix $\proj$ must have certain properties in order
to be suitable for CS. One desirable property (which leads to the
theoretical results we mention in Section~\ref{sec:csresults}) is
known as the Restricted Isometry Property
(RIP)~\cite{CandesUES,CandesECLP,CandesSSR}. We say a matrix $\proj$
meets the {\em RIP of order $\sparsity$ with respect to the basis
$\Psi$} if for some $\delta_\sparsity > 0$,
$$
(1-\delta_\sparsity) \norm{\alpha} \le \norm{\proj \Psi \alpha} \le
(1+\delta_\sparsity) \norm{\alpha}
$$
holds for all $\alpha \in \mathbb{R} ^\dim$ with $\|\alpha\|_0 \le
\sparsity$. Intuitively, the RIP can be viewed as guaranteeing a
{\em stable embedding} of the collection of $\sparsity$-sparse
signals within the measurement space $ \mathbb{R} ^\pdim$. In particular,
supposing the RIP of order $2\sparsity$ is satisfied with respect to
the basis $\Psi$, then for all pairs of $\sparsity$-sparse signals
$x_1, x_2 \in \mathbb{R} ^\dim$, we have
\begin{equation}
(1-\delta_{2\sparsity}) \norm{x_1-x_2} \le \norm{\proj x_1 - \proj
x_2} \le (1+\delta_{2\sparsity}) \norm{x_1-x_2}. \label{eq:rip2}
\end{equation}
Although deterministic constructions of matrices meeting the RIP are
still a work in progress, it is known that the RIP often be met by
choosing $\proj$ randomly from an acceptable distribution. For
example, let $\Psi$ be a fixed orthonormal basis for $ \mathbb{R} ^\dim$
and suppose that
\begin{equation}
\pdim \ge C_0 \sparsity
\log(\dim/\sparsity) \label{eq:nummeas}
\end{equation}
for some constant $C_0$. Then supposing that the entries of
the $\pdim \times \dim$ matrix $\proj$ are drawn as independent,
identically distributed Gaussian random variables with mean $0$ and
variance $\frac{1}{\pdim}$, it follows that with high probability
$\proj$ meets the RIP of order $\sparsity$ with respect to the basis
$\Psi$. Two aspects of this construction deserve special notice:
first, the number $\pdim$ of measurements required is linearly
proportional to the information level $\sparsity$, and second,
neither the sparse basis $\Psi$ nor the locations of the nonzero
entries of $\alpha$ need be known when designing the measurement
operator $\proj$. Other random distributions for $\proj$ may also be
used, all requiring approximately the same number of measurements.
One of these distributions~\cite{dasgupta99elementary,JLCS} dictates
that $\proj = \sqrt{\dim/\pdim}\, \Xi$, where $\Xi$ is an $\pdim
\times \dim$ matrix having orthonormal rows that span a random
$\pdim$-dimensional subspace of $ \mathbb{R} ^\dim$. We refer to such
choice of $\proj$ as a {\em random orthoprojector}.\footnote{Our
previous use of the term ``random orthoprojector'' in~\cite{mbwFocm}
excluded the normalization factor of $\sqrt{\dim/\pdim}$. However we
find it more appropriate to include this factor in the current
paper.}
\subsection{Signal recovery and sketching}
\label{sec:csresults}
Although the sparse structure of a signal $x$ need not be known when
collecting measurements $y = \proj x$, a hallmark of CS is the use
of the sparse model in order to facilitate understanding from the
compressive measurements. A variety of algorithms have been proposed
to answer Q1 (signal recovery), where we seek to solve the
apparently undercomplete set of $\pdim$ linear equations $y = \proj
x$ for $\dim$ unknowns. The canonical
method~\cite{DonohoCS,CandesUES,CandesRUP} is known as {\em
$\ell_1$-minimization} and is formulated as follows: first solve
\begin{equation}
\widehat{\alpha} = \arg\min_{\alpha' \in \mathbb{R} ^\dim} \|\alpha'\|_1
~\mathrm{subject~to}~ y = \proj \Psi \alpha', \label{eq:l1min}
\end{equation}
and then set $\widehat{x} = \Psi \widehat{\alpha}$. Under this
recovery program, the following bounds are known.
\begin{thm}{\em \cite{candes2008rip}}
Suppose that $\proj$ satisfies the RIP of order $2\sparsity$ with
respect to $\Psi$ and with constant $\delta_{2\sparsity} <
\sqrt{2}-1$. Let $x \in \mathbb{R} ^\dim$, suppose $y = \proj x$, and let
the recovered estimates $\widehat{\alpha}$ and $\widehat{x}$ be as
defined above. Then
\begin{equation}
\norm{x-\widehat{x}} = \norm{\alpha-\widehat{\alpha}} \le C_1
\sparsity^{-1/2} \|\alpha-\alpha_\sparsity\|_1 \label{eq:mixed1}
\end{equation}
for a constant $C_1$. In particular, if $x$ is
$\sparsity$-sparse, then $\widehat{x} = x$. \label{theo:mixed1}
\end{thm}
This result can be extended to account for measurement noise.
\begin{thm}{\em \cite{candes2008rip}} Suppose that $\proj$
satisfies the RIP of order $2\sparsity$ with respect to $\Psi$ and
with constant $\delta_{2\sparsity} < \sqrt{2}-1$. Let $x \in
\mathbb{R} ^\dim$, and suppose that
$$
y = \proj x + \noise
$$
where $\norm{\noise} \le \epsilon$. Then let
$$ \widehat{\alpha} =
\arg\min_{\alpha' \in \mathbb{R} ^\dim} \|\alpha'\|_1
~\mathrm{subject~to}~ \norm{y - \proj \Psi \alpha'} \le \epsilon,
$$
and set $\widehat{x} = \Psi \widehat{\alpha}$. Then
\begin{equation}
\norm{x-\widehat{x}} = \norm{\alpha-\widehat{\alpha}} \le C_1
\sparsity^{-1/2} \|\alpha-\alpha_\sparsity\|_1 + C_2
\epsilon. \label{eq:mixed2}
\end{equation}
for constants $C_1$ (which is the same as above) and
$C_2$.
\label{theo:mixed2}
\end{thm}
These results are not unique to $\ell_1$ minimization; similar
bounds have been established for signal recovery using greedy
iterative algorithms ROMP~\cite{romp} and CoSAMP~\cite{cosamp}.
Bounds of this type are extremely encouraging for signal processing.
From only $\pdim$ measurements, it is possible to recover $x$ with
quality that is comparable to its proximity to the nearest
$\sparsity$-sparse signal, and if $x$ itself is $\sparsity$-sparse
and there is no measurement noise, then $x$ can be recovered
exactly. Moreover, despite the apparent ill-conditioning of the
inverse problem, the measurement noise is not dramatically amplified
in the recovery process.
These bounds are known as {\em deterministic}, {\em
instance-optimal} bounds because they hold deterministically for any
$\proj$ that meets the RIP, and because for a given $\proj$ they
give a guarantee for recovery of any $x \in \mathbb{R} ^\dim$ based on its
proximity to the concise model.
The use of $\ell_1$ as a measure for proximity to the concise model
(on the right hand side of (\ref{eq:mixed1}) and (\ref{eq:mixed2}))
arises due to the difficulty in establishing $\ell_2$ bounds on the
right hand side. Indeed, it is known that deterministic $\ell_2$
instance-optimal bounds cannot exist that are comparable to
(\ref{eq:mixed1}) and (\ref{eq:mixed2}). In particular, for any
$\proj$, to ensure that $\norm{x-\widehat{x}} \le C_3
\norm{x-x_\sparsity}$ for all $x$, it is known~\cite{CDD} that this
requires that $\pdim \ge C_4 \dim$ regardless of $\sparsity$.
However, it is possible to obtain an instance-optimal $\ell_2$ bound
for sparse signal recovery in the noise-free setting by changing
from a deterministic formulation to a {\em probabilistic}
one~\cite{CDD,devore08in}. In particular, by considering any given
$x \in \mathbb{R} ^\dim$, it is possible to show that for {\em most}
random $\proj$, letting the measurements $y = \proj x$, and
recovering $\widehat{x}$ via $\ell_1$-minimization (\ref{eq:l1min}),
it holds that
\begin{equation}
\norm{x-\widehat{x}} \le C_5
\norm{x-x_\sparsity}.\label{eq:devore}
\end{equation}
While the proof of this statement~\cite{devore08in} does not involve
the RIP directly, it holds for many of the same random distributions
that work for RIP matrices, and it requires the same number of
measurements (\ref{eq:nummeas}) up to a constant.
Similar bounds hold for the closely related problem of Q2
(sketching), where the goal is to use the compressive measurement
vector $y$ to identify and report only approximately $\sparsity$
expansion coefficients that best describe the original signal, i.e.,
a sparse approximation to $\alpha_\sparsity$. In the case where
$\Psi = I$, an efficient randomized measurement process coupled with
a customized recovery algorithm~\cite{gilbert2007osa} provides
signal sketches that meet a deterministic mixed-norm $\ell_2/\ell_1$
instance-optimal bound analogous to~(\ref{eq:mixed1}). A desirable
aspect of this construction is that the computational complexity
scales with only $\log(\dim)$ (and is polynomial in $\sparsity$);
this is possible because only approximately $\sparsity$ pieces of
information must be computed to describe the signal. For signals
that are sparse in the Fourier domain ($\Psi$ consists of the DFT
vectors), probabilistic $\ell_2/\ell_2$ instance-optimal bounds have
also been established~\cite{gilbert05im} that are analogous to
(\ref{eq:devore}).
\section{Compressive Measurements of Manifold-Modeled Signals}
\label{sec:manifolds}
\subsection{Manifold models}
\label{sec:mmodels}
As we have discussed in Section~\ref{sec:mmodelsunder}, there are
many possible modeling frameworks for capturing concise signal
structure. Among these possibilities are the broad class of manifold
models.
Manifold models arise, for example, in settings where the signals of
interest vary continuously as a function of some $\mdim$-dimensional
parameter. Suppose, for instance, that there exists some parameter
$\theta$ that controls the generation of the signal. We let
$x_\theta \in \mathbb{R} ^\dim$ denote the signal corresponding to the
parameter $\theta$, and we let $\Theta$ denote the
$\mdim$-dimensional parameter space from which $\theta$ is drawn. In
general, $\Theta$ itself may be a $\mdim$-dimensional manifold and
need not be embedded in an ambient Euclidean space. For example,
supposing $\theta$ describes the 1-D rotation parameter in a
top-down satellite image, we have $\Theta = S^1$.
Under certain conditions on the parameterization $\theta \mapsto
x_\theta$, it follows that
$$
\manifold := \{x_\theta: \theta \in \Theta\}
$$
forms a $\mdim$-dimensional {\em submanifold} of $ \mathbb{R} ^\dim$. An
appropriate visualization is that the set $\manifold$ forms a
nonlinear $\mdim$-dimensional ``surface'' within the
high-dimensional ambient signal space $ \mathbb{R} ^\dim$. Depending on the
circumstances, we may measure the distance between points two points
$x_{\theta_1}$ and $x_{\theta_2}$ on the manifold $\manifold$ using
either the ambient Euclidean distance
$$
\norm{x_{\theta_1}-x_{\theta_2}}
$$
or the geodesic distance along the manifold, which we denote as
$\gdist(x_{\theta_1},x_{\theta_2})$. In the case where the geodesic
distance along $\manifold$ equals the native distance in parameter
space, i.e., when
\begin{equation}
\gdist(x_{\theta_1},x_{\theta_2}) = d_\Theta(\theta_1, \theta_2),
\label{eq:isometry}
\end{equation}
we say that $\manifold$ is {\em isometric} to $\Theta$. The
definition of the distance $d_\Theta(\theta_1, \theta_2)$ depends on
the appropriate metric for the parameter space $\Theta$; supposing
$\Theta$ is a convex subset of Euclidean space, then we have
$d_\Theta(\theta_1, \theta_2) = \norm{\theta_1-\theta_2}$.
While our discussion above concentrates on the case of manifolds
$\manifold$ generated by underlying parameterizations, we stress
that manifolds have also been proposed as approximate
low-dimensional models within $ \mathbb{R} ^\dim$ for nonparametric signal
classes such as images of human faces or handwritten
digits~\cite{Eigenfaces,digits,broomhead01wh}. These signal families
may also be considered.
The results we present in this paper will make reference to certain
characteristic properties of the manifold under study. These terms
are originally defined in~\cite{niyogi,mbwFocm} and are repeated
here for completeness. First, our results will depend on a measure
of regularity for the manifold. For this purpose, we adopt the {\em
condition number} defined recently by Niyogi et al.~\cite{niyogi}.
\begin{definition}{\em \cite{niyogi}}
Let $\manifold$ be a compact Riemannian submanifold of $ \mathbb{R} ^\dim$.
The {\em condition number} is defined as $1/\condition$, where
$\condition$ is the largest number having the following property:
The open normal bundle about $\manifold$ of radius $r$ is embedded
in $ \mathbb{R} ^\dim$ for all $r < \condition$. \label{def:cn}
\end{definition}
The condition number $1/\condition$ controls both local properties
and global properties of the manifold. Its role is summarized in two
key relationships~\cite{niyogi}. First, the the curvature of any
unit-speed geodesic path on $\manifold$ is bounded by
$1/\condition$. Second, at long geodesic distances, the condition
number controls how close the manifold may curve back upon itself.
For example, supposing $x_1,x_2 \in \manifold$ with $\gdist(x_1,x_2)
> \condition$, it must hold that $\norm{x_1-x_2} > \condition/2$.
We also require a notion of ``geodesic covering regularity'' for a
manifold. While this property is not the focus of the present paper,
we include its definition in Appendix~\ref{app:reg} for
completeness.
We conclude with a brief but concrete example to illustrate specific
values for these quantities. Let $\dim
> 0$, $\kappa > 0$, $\Theta = \mathbb{R} \!\! \mod 2\pi$, and suppose $x_\theta
\in \mathbb{R} ^\dim$ is given by
$$
x_\theta = [\kappa \cos(\theta); ~ \kappa \sin(\theta); ~ 0; ~0;
\cdots 0]^T.
$$
In this case, $\manifold = \{x_\theta: \theta \in \Theta\}$ forms a
circle of radius $\kappa$ in the $x(1), x(2)$ plane. The manifold
dimension $\mdim=1$, the condition number $\condition = \kappa$, and
the geodesic covering regularity $\regularity$ can be chosen as any
number larger than $\frac{1}{2}$. We also refer in our results to
the $\mdim$-dimensional volume $\volume$ of the $\manifold$, which
in this example corresponds to the circumference $2\pi \kappa$ of
the circle.
\subsection{Stable embeddings of manifolds}
\label{sec:mstable}
In cases where the signal class of interest $\manifold$ forms a
low-dimensional submanifold of $ \mathbb{R} ^\dim$, we have theoretical
justification that the information necessary to distinguish and
recover signals $x \in \manifold$ can be well-preserved under a
sufficient number of compressive measurements $y = \proj x$. In
particular, we have recently shown that an RIP-like property holds
for families of manifold-modeled signals.
\begin{thm} {\em \cite{mbwFocm}}
Let $\manifold$ be a compact $\mdim$-dimensional Riemannian
submanifold of $ \mathbb{R} ^\dim$ having condition number $1/\condition$,
volume $\volume$, and geodesic covering regularity $R$. Fix $0 <
\epsilon < 1$ and $0 < \rho < 1$. Let $\proj$ be a random $\pdim \times
\dim$ orthoprojector with
\begin{equation} \pdim = O\left(\frac{\mdim \log(\dim \volume \regularity \condition^{-1} \epsilon^{-1}) \log(1/\rho)}{\epsilon^2}
\right).
\label{eq:mmeasmain}
\end{equation}
If $\pdim \le \dim$, then with probability at least $1-\rho$ the
following statement holds: For every pair of points $x_1,x_2 \in
\manifold$,
\begin{equation}
(1-\epsilon) \norm{x_1-x_2} \le \norm{\proj x_1 - \proj x_2} \le
(1+\epsilon) \norm{x_1-x_2}. \label{eq:mmain}
\end{equation}
\label{theo:manifoldjl}
\end{thm}
The proof of this theorem involves the Johnson-Lindenstrauss
Lemma~\cite{dasgupta99elementary,dfrp,JLCS}, which guarantees a
stable embedding for a finite point cloud under a sufficient number
of random projections. In essence, manifolds with higher volume or
with greater curvature have more complexity and require a more dense
covering for application of the Johnson-Lindenstrauss Lemma; this
leads to an increased number of measurements~(\ref{eq:mmeasmain}).
By comparing (\ref{eq:rip2}) with (\ref{eq:mmain}), we see a strong
analogy to the RIP of order $2\sparsity$. This theorem establishes
that, like the class of $\sparsity$-sparse signals, a collection of
signals described by a $\mdim$-dimensional manifold $\manifold
\subset \mathbb{R} ^\dim$ can have a stable embedding in an
$\pdim$-dimensional measurement space. Moreover, the requisite
number of random measurements $\pdim$ is once again linearly
proportional to the information level (or number of degrees of
freedom) $\mdim$.
As was the case with the RIP for sparse signal processing, this
result has a number of possible implications for manifold-based
signal processing. Individual signals obeying a manifold model can
be acquired and stored efficiently using compressive measurements,
and it is unnecessary to employ the manifold model itself as part of
the compression process. Rather, the model need be used only for
signal understanding from the compressive measurements. Problems
such as Q1 (signal recovery) and Q3 (parameter estimation) can be
addressed. We have reported promising experimental results with
various classes of parametric
signals~\cite{mbwPhdThesis,davenport2007sfc}. We have also extended
Theorem~\ref{theo:manifoldjl} to the case of multiple manifolds that
are simultaneously embedded~\cite{davenport2007sfc}; this allows
both the classification of an observed object to one of several
possible models (different manifolds) and the estimation of a
parameter within that class (position on a manifold). Moreover,
collections of signals obeying a manifold model (such as multiple
images of a scene photographed from different perspectives) can be
acquired using compressive measurements, and the resulting manifold
structure will be preserved among the suite of measurement vectors
in $ \mathbb{R} ^\pdim$. We have provided empirical and theoretical support
for the use of manifold learning in the reduced-dimensional
space~\cite{hegde2007rpm}; this can dramatically simplify the
computational and storage demands on a system for processing large
databases of signals.
\subsection{Signal recovery and parameter estimation}
\label{sec:problem}
In this paper, we provide theoretical justification for the
encouraging experimental results that have been observed for
problems Q1 (signal recovery) and Q3 (parameter estimation).
To be specific, let us consider a length-$\dim$ signal $x$ that,
rather than being $\sparsity$-sparse, we assume lives on or near
some known $\mdim$-dimensional manifold $\manifold \subset
\mathbb{R} ^\dim$. From a collection of measurements
$$y = \proj x + \noise,$$
where $\proj$ is a random $\pdim \times \dim$ matrix and $\noise \in
\mathbb{R} ^\pdim$ is an additive noise vector, we would like to recover
either $x$ or a parameter $\theta$ that generates $x$.
For the signal recovery problem, we will consider the following as a
method for estimating $x$:
\begin{equation}
\widehat{x} = \arg\min_{x' \in \manifold} \norm{y-\proj
x'},\label{eq:csmrecover}
\end{equation}
supposing here and elsewhere that the minimum is uniquely defined.
We also let $x^\ast$ be the optimal ``nearest neighbor'' to $x$ on
$\manifold$, i.e.,
\begin{equation}
x^\ast = \arg\min_{x' \in \manifold} \norm{x-x'}. \label{eq:csmopt}
\end{equation}
To consider signal recovery successful, we would like to guarantee
that $\norm{x-\widehat{x}}$ is not much larger than
$\norm{x-x^\ast}$.
For the parameter estimation problem, where we presume $x \approx
x_\theta$ for some $\theta \in \Theta$, we propose a similar method
for estimating $\theta$ from the compressive measurements:
\begin{equation}
\widehat{\theta} = \arg\min_{\theta' \in \Theta} \norm{y-\proj
x_{\theta'}}.\label{eq:parmrecover}
\end{equation}
Let $\theta^\ast$ be the ``optimal estimate'' that could be obtained
using the full data $x \in \mathbb{R} ^\dim$, i.e.,
\begin{equation}
\theta^\ast = \arg\min_{\theta' \in \Theta} \norm{x-x_{\theta'}}.
\label{eq:parmopt}
\end{equation}
(If $x = x_\theta$ exactly for some $\theta$, then $\theta^\ast =
\theta$; otherwise this formulation allows us to consider signals
$x$ that are not precisely on the manifold $\manifold$ in
$ \mathbb{R} ^\dim$. This generalization has practical relevance; a local
image block, for example, may only approximately resemble a straight
edge, which has a simple parameterization.) To consider parameter
estimation successful, we would like to guarantee that
$d_\Theta(\widehat{\theta},\theta^\ast)$ is small.
As we will see, bounds pertaining to accurate signal recovery can
often be extended to imply accurate parameter estimation as well.
However, the relationships between distance $d_\Theta$ in parameter
space and distances $d_\manifold$ and $\|\cdot\|_2$ in the signal
space can vary depending on the parametric signal model under study.
Thus, for the parameter estimation problem, our ability to provide
generic bounds on $d_\Theta(\widehat{\theta},\theta^\ast)$ will be
restricted. In this paper we focus primarily on the signal recovery
problem and provide preliminary results for the parameter estimation
problem that pertain most strongly to the case of isometric
parameterizations.
In this paper, we do not confront in depth the question of how a
recovery program such as (\ref{eq:csmrecover}) can be efficiently
solved. Some discussion of this matter is provided
in~\cite{mbwFocm}, with application-specific examples provided
in~\cite{mbwPhdThesis,davenport2007sfc}. Unfortunately, it is
difficult to propose a single general-purpose algorithm for solving
(\ref{eq:csmrecover}) in $ \mathbb{R} ^\pdim$, as even the problem
(\ref{eq:csmopt}) in $ \mathbb{R} ^\dim$ may be difficult to solve
depending on certain nuances (such as topology) of the individual
manifold. Nonetheless, iterative algorithms such as Newton's
method~\cite{WakinSPIEiam} have proved helpful in many problems to
date. Additional complications arise when the manifold $\manifold$
is non-differentiable, as may happen when the signals $x$ represent
2-D images. However, just as a multiscale regularization can be
incorporated into Newton's method for solving (\ref{eq:csmopt})
(see~\cite{WakinSPIEiam}), an analogous regularization can be
incorporated into a compressive measurement operator $\proj$ to
facilitate Newton's method for solving (\ref{eq:csmrecover})
(see~\cite{donohoECS,mbwPhdThesis}). For manifolds that lack
differentiability, additional care must be taken when applying
results such as Theorem~\ref{theo:manifoldjl}; we defer a study of
these matters to a subsequent paper.
In the following, we will consider both deterministic and
probabilistic instance-optimal bounds for signal recovery and
parameter estimation, and we will draw comparisons to the
sparsity-based CS results of Section~\ref{sec:csresults}. Our bounds
are formulated in terms of generic properties of the manifold (as
mentioned in Section~\ref{sec:mmodels}), which will vary from signal
model to signal model. In some cases, calculating these may be
possible, whereas in other cases it may not. Nonetheless, we feel
the results in this paper highlight the relative importance of these
properties in determining the requisite number of measurements.
Finally, to simplify analysis we will focus on random
orthoprojectors for the measurement operator $\proj$, although our
results may be extended to other random distributions such as the
Gaussian~\cite{mbwFocm}.
\section{A deterministic instance-optimal bound in $\ell_2$}
\label{sec:deter}
We begin by seeking a deterministic instance-optimal bound. That is,
for a measurement matrix $\proj$ that meets (\ref{eq:mmain}) for all
$x_1, x_2 \in \manifold$, we seek an upper bound for the relative
reconstruction error
$$
\frac{\norm{x-\widehat{x}}}{\norm{x-x^\ast}}
$$
that holds uniformly for all $x \in \mathbb{R} ^\dim$. In this section we
consider only the signal recovery problem; however, similar bounds
would apply to parameter estimation. We have the following result
for the noise-free case, which applies not only to the manifolds
described in Theorem~\ref{theo:manifoldjl} but also to more general
sets.
\begin{thm}
Let $\manifold \subset \mathbb{R} ^\dim$ be any subset of $ \mathbb{R} ^\dim$,
and let $\proj$ denote an $\pdim \times \dim$ orthoprojector
satisfying (\ref{eq:mmain}) for all $x_1,x_2 \in \manifold$. Suppose
$x \in \mathbb{R} ^\dim$, let $y = \proj x$, and let the recovered
estimate $\widehat{x}$ and the optimal estimate $x^\ast$ be as
defined in (\ref{eq:csmrecover}) and (\ref{eq:csmopt}). Then
\begin{equation}
\frac{\norm{x-\widehat{x}}}{\norm{x-x^\ast}} \le
\sqrt{\frac{4\dim}{\pdim (1-\epsilon)^2} - 3 +
2\sqrt{\frac{\dim}{\pdim(1-\epsilon)^2}-1}}. \label{eq:kappabound}
\end{equation}
\label{thm:kappa}
\end{thm}
\noindent {\bf Proof:} See Appendix~\ref{app:kappa}.
~
As $\frac{\pdim}{\dim} \rightarrow 0$, the bound on the right hand side
of (\ref{eq:kappabound}) grows as
$\frac{2}{1-\epsilon}\sqrt{\frac{\dim}{\pdim}}$. Unfortunately, this
is not desirable for signal recovery. Supposing, for example, that
we wish to ensure $\norm{x-\widehat{x}} \le C_6
\norm{x-x^\ast}$ for all $x \in \mathbb{R} ^\dim$, then using the bound
(\ref{eq:kappabound}) we would require that $\pdim \ge C_7
\dim$ regardless of the dimension $\mdim$ of the manifold.
The weakness of this bound is a geometric necessity; indeed, the
bound itself is quite tight in general, as the following simple
example illustrates. Suppose $\dim \ge 2$ and let $\manifold$ denote
the line segment in $ \mathbb{R} ^\dim$ joining the points $(0,0,\dots,0)$
and $(1,0,0,\dots,0)$. Let $0 \le \gamma < \pi/2$ for some $\gamma$,
let $\pdim = 1$, and let the $1 \times \dim$ measurement matrix
$$
\proj = \sqrt{\dim} \, [\cos(\gamma); ~ -\!\sin(\gamma); ~ 0; ~ 0; ~
\cdots; ~0].
$$
Any $x_1 \in \manifold$ we may write as $x_1 =
(x_1(1),0,0,\dots,0)$, and it follows that $\proj x_1 = \sqrt{\dim}
\cos(\gamma)x_1(1)$. Thus for any pair $x_1,x_2 \in \manifold$, we
have
$$
\frac{\norm{\proj x_1 -\proj x_2}}{\norm{x_1 - x_2}} =
\frac{|\sqrt{\dim}\cos(\gamma)x_1(1)-\sqrt{\dim}\cos(\gamma)x_2(1)|}{|x_1(1)-x_2(1)|}
= \sqrt{\dim}\cos(\gamma).
$$
We suppose that $\cos(\gamma) < \frac{1}{\sqrt{\dim}}$ and thus
referring to equation (\ref{eq:mmain}) we have $(1-\epsilon)
=\sqrt{\dim}\cos(\gamma)$. Now, we may consider the signal $x =
(1,\tan(\pi/2-\gamma),0,0,\dots,0)$. We then have that $x^\ast =
(1,0,0,\dots,0)$, and $\norm{x-x^\ast} = \tan(\pi/2-\gamma)$. We
also have that $\proj x =
\sqrt{\dim}(\cos(\gamma)-\sin(\gamma)\tan(\pi/2-\gamma)) = 0$. Thus
$\widehat{x} = (0,0,\dots,0)$ and $\norm{x-\widehat{x}} =
\frac{1}{\cos(\pi/2-\gamma)}$, and so
$$
\frac{\norm{x-\widehat{x}}}{\norm{x-x^\ast}} =
\frac{1}{\cos(\pi/2-\gamma)\tan(\pi/2-\gamma)} =
\frac{1}{\sin(\pi/2-\gamma)} = \frac{1}{\cos(\gamma)} =
\frac{\sqrt{\dim}}{1-\epsilon}.
$$
It is worth recalling that, as we discussed in
Section~\ref{sec:csresults}, similar difficulties arise in
sparsity-based CS when attempting to establish a deterministic
$\ell_2$ instance-optimal bound. In particular, to ensure that
$\norm{x-\widehat{x}} \le C_3 \norm{x-x_\sparsity}$ for all $x
\in \mathbb{R} ^\dim$, it is known~\cite{CDD} that this requires $\pdim
\ge C_4 \dim$ regardless of the sparsity level $\sparsity$.
In sparsity-based CS, there have been at least two types of
alternative approaches. The first are the deterministic
``mixed-norm'' results of the type given in (\ref{eq:mixed1}) and
(\ref{eq:mixed2}). These involve the use of an alternative norm such
as the $\ell_1$ norm to measure the distance from the coefficient
vector $\alpha$ to its best $\sparsity$-term approximation
$\alpha_\sparsity$. While it may be possible to pursue similar
directions for manifold-modeled signals, we feel this is undesirable
as a general approach because when sparsity is no longer part of the
modeling framework, the $\ell_1$ norm has less of a natural meaning.
Instead, we prefer to seek bounds using $\ell_2$, as that is the
most conventional norm used in signal processing to measure energy
and error.
Thus, the second type of alternative bounds in sparsity-based CS
have involved $\ell_2$ bounds in probability, as we discussed in
Section~\ref{sec:csresults}. Indeed, the performance of both
sparsity-based and manifold-based CS is often much better in
practice than a deterministic $\ell_2$ instance-optimal bound might
indicate. The reason is that, for any $\proj$, such bounds consider
the {\em worst case} signal over all possible $x \in \mathbb{R} ^\dim$.
Fortunately, this worst case is not typical. As a result, it is
possible to derive much stronger results that consider any given
signal $x \in \mathbb{R} ^\dim$ and establish that for most random
$\proj$, the recovery error of that signal $x$ will be small.
\section{Probabilistic instance-optimal bounds in $\ell_2$}
\label{sec:prob}
For a given measurement operator $\proj$, our bound in
Theorem~\ref{thm:kappa} applies uniformly to any signal in
$ \mathbb{R} ^\dim$. However, a much sharper bound can be obtained by
relaxing the deterministic requirement.
\subsection{Signal recovery}
Our first bound applies to the signal recovery problem, and we
include the consideration of additive noise in the measurements.
\begin{thm} Suppose $x \in \mathbb{R} ^\dim$. Let
$\manifold$ be a compact $\mdim$-dimensional Riemannian submanifold
of $ \mathbb{R} ^\dim$ having condition number $1/\condition$, volume
$\volume$, and geodesic covering regularity $R$. Fix $0 < \epsilon < 1$
and $0 < \rho < 1$. Let $\proj$ be a random $\pdim \times \dim$
orthoprojector, chosen independently of $x$, with
\begin{equation}
\pdim = O\left(\frac{\mdim \log(\dim \volume \regularity
\condition^{-1} \epsilon^{-1}) \log(1/\rho)}{\epsilon^2} \right).
\label{eq:bound3meas}
\end{equation}
Let $\noise \in \mathbb{R} ^\pdim$, let $y = \proj x + \noise$, and let
the recovered estimate $\widehat{x}$ and the optimal estimate
$x^\ast$ be as defined in (\ref{eq:csmrecover}) and
(\ref{eq:csmopt}). If $\pdim \le \dim$, then with probability at
least $1-\rho$ the following statement holds:
\begin{equation}
\norm{x-\widehat{x}} \le (1 + 0.25\epsilon)\norm{x-x^\ast} +
(2+0.32\epsilon)\norm{\noise} + \frac{\epsilon^2 \condition}{936
\dim}.\label{eq:bound3}
\end{equation}
\label{theo:bound3}
\end{thm}
\noindent {\bf Proof:} See Appendix~\ref{app:bound3}.
~
The proof of this theorem, like that of
Theorem~\ref{theo:manifoldjl}, involves the Johnson-Lindenstrauss
Lemma. Our proof of Theorem~\ref{theo:bound3} extends the proof of
Theorem~\ref{theo:manifoldjl} by adding the points $x$ and $x^\ast$
to the finite sampling of points drawn from $\manifold$ that are
used to establish (\ref{eq:mmain}).
Let us now compare and contrast our bound with the analogous results
for sparsity-based CS. Like Theorem~\ref{theo:mixed2}, we consider
the problem of signal recovery in the presence of additive
measurement noise. Both bounds relate the recovery error
$\norm{x-\widehat{x}}$ to the proximity of $x$ to its nearest
neighbor in the concise model class (either $x_\sparsity$ or
$x^\ast$ depending on the model), and both bounds relate the
recovery error $\norm{x-\widehat{x}}$ to the amount $\norm{\noise}$
of additive measurement noise.
However, Theorem~\ref{theo:mixed2} is a deterministic bound whereas
Theorem~\ref{theo:bound3} is probabilistic, and our bound
(\ref{eq:bound3}) measures proximity to the concise model in the
$\ell_2$ norm, whereas (\ref{eq:mixed2}) uses the $\ell_1$ norm.
Our bound can also be compared with (\ref{eq:devore}), as both are
instance-optimal bounds in probability, and both use the $\ell_2$
norm to measure proximity to the concise model. However, we note
that unlike (\ref{eq:devore}), our bound (\ref{eq:bound3}) allows
the consideration of measurement noise.
Finally, we note that there is an additional term $\frac{\epsilon^2
\condition}{936 \dim}$ appearing on the right hand side of
(\ref{eq:bound3}). This term becomes relevant only when both
$\norm{x-x^\ast}$ and $\norm{\noise}$ are significantly smaller than
the condition number $\condition$, since $\epsilon^2 < 1$ and
$\frac{1}{936 \dim} \ll 1$. Indeed, in these regimes the signal
recovery remains accurate (much smaller than $\condition$), but the
quantity $\norm{x-\widehat{x}}$ may not remain strictly proportional
to $\norm{x-x^\ast}$ and $\norm{\noise}$. The bound may also be
sharpened by artificially assuming a condition number $1/\condition'
> 1/\condition$ for the purpose of choosing a number of measurements
$\pdim$ in (\ref{eq:bound3meas}). This will decrease the last term
in (\ref{eq:bound3}) as $\frac{\epsilon^2 \condition'}{936 \dim}$. In
the case where $\noise = 0$, it is also possible to resort to the
bound (\ref{eq:kappabound}); this bound is inferior to
(\ref{eq:bound3}) when $\norm{x-x^\ast}$ is large but ensures that
$\norm{x-\widehat{x}} \rightarrow 0$ when $\norm{x-x^\ast}
\rightarrow 0$.
\subsection{Parameter estimation}
Above we have derived a bound for the signal recovery problem, with
an error metric that measures the discrepancy between the recovered
signal $\widehat{x}$ and the original signal $x$.
However, in some applications it may be the case that the original
signal $x \approx x_{\theta^\ast}$, where $\theta^\ast \in \Theta$
is a parameter of interest. In this case we may be interested in
using the compressive measurements $y = \proj x + \noise$ to solve
the problem (\ref{eq:parmrecover}) and recover an estimate
$\widehat{\theta}$ of the underlying parameter.
Of course, these two problems are closely related. However, we
should emphasize that guaranteeing $\norm{x-\widehat{x}} \approx
\norm{x-x^\ast}$ does not automatically guarantee that
$\gdist(x_{\widehat{\theta}},x_{\theta^\ast})$ is small (and
therefore does not ensure that
$d_\Theta(\widehat{\theta},\theta^\ast)$ is small). If the manifold
is shaped like a horseshoe, for example, then it could be the case
that $x_{\theta^\ast}$ sits at the end of one arm but
$x_{\widehat{\theta}}$ sits at the end of the opposing arm. These
two points would be much closer in a Euclidean metric than in a
geodesic one.
Consequently, in order to establish bounds relevant for parameter
estimation, our concern focuses on guaranteeing that the geodesic
distance $\gdist(x_{\widehat{\theta}},x_{\theta^\ast})$ is itself
small.
\begin{thm}
Suppose $x \in \mathbb{R} ^\dim$. Let $\manifold$ be a compact
$\mdim$-dimensional Riemannian submanifold of $ \mathbb{R} ^\dim$ having
condition number $1/\condition$, volume $\volume$, and geodesic
covering regularity $R$. Fix $0 < \epsilon < 1$ and $0 < \rho < 1$. Let
$\proj$ be a random $\pdim \times \dim$ orthoprojector, chosen
independently of $x$, with
\begin{equation*}
\pdim = O\left(\frac{\mdim \log(\dim \volume \regularity
\condition^{-1} \epsilon^{-1}) \log(1/\rho)}{\epsilon^2} \right).
\end{equation*}
Let $\noise \in \mathbb{R} ^\pdim$, let $y = \proj x + \noise$, and let
the recovered estimate $\widehat{x}$ and the optimal estimate
$x^\ast$ be as defined in (\ref{eq:csmrecover}) and
(\ref{eq:csmopt}). If $\pdim \le \dim$ and if $1.16\norm{\noise} +
\norm{x-x^\ast} \le \condition/5$, then with probability at least
$1-\rho$ the following statement holds:
\begin{equation}
\gdist(\widehat{x}, x^\ast) \le (4 + 0.5\epsilon)\norm{x-x^\ast}
+ (4+0.64\epsilon)\norm{\noise} + \frac{\epsilon^2 \condition}{468 \dim}.
\label{eq:bound4}
\end{equation}
\label{theo:bound4}
\end{thm}
\noindent {\bf Proof:} See Appendix~\ref{app:bound4}.
~
In several ways, this bound is similar to (\ref{eq:bound3}). Both
bounds relate the recovery error to the proximity of $x$ to its
nearest neighbor $x^\ast$ on the manifold and to the amount
$\norm{\noise}$ of additive measurement noise.
Both bounds also have an additive term on the right hand side that
is small in relation to the condition number $\tau$.
In contrast, (\ref{eq:bound4}) guarantees that the recovered
estimate $\widehat{x}$ is near to the optimal estimate $x^\ast$ in
terms of geodesic distance along the manifold. Establishing this
condition required the additional assumption that $1.16\norm{\noise}
+ \norm{x-x^\ast} \le \condition/5$. Because $\condition$ relates to
the degree to which the manifold can curve back upon itself at long
geodesic distances, this assumption prevents exactly the type of
``horseshoe'' problem that was mentioned above, where it may happen
that $\gdist(\widehat{x}, x^\ast) \gg \norm{\widehat{x}-x^\ast}$.
Suppose, for example, it were to happen that $\norm{x-x^\ast}
\approx \condition$ and $x$ was approximately equidistant from both
ends of the horseshoe; a small distortion of distances under $\proj$
could then lead to an estimate $\widehat{x}$ for which
$\norm{x-\widehat{x}} \approx \norm{x-x^\ast}$ but
$\gdist(\widehat{x},x^\ast) \gg 0$. Similarly, additive noise could
cause a similar problem of ``crossing over'' in the measurement
space. Although our bound provides no guarantee in these situations,
we stress that under these circumstances, accurate parameter
estimation would be difficult (or perhaps even unimportant) in the
original signal space $ \mathbb{R} ^\dim$.
Finally, we revisit the situation where the original signal $x
\approx x_{\theta^\ast}$ for some $\theta^\ast \in \Theta$ (with
$\theta^\ast$ satisfying (\ref{eq:parmopt})), where the measurements
$y = \proj x + \noise$, and where the recovered estimate
$\widehat{\theta}$ satisfies (\ref{eq:parmrecover}). We consider the
question of whether (\ref{eq:bound4}) can be translated into a bound
on $d_\Theta(\widehat{\theta},\theta^\ast)$. As described in
Section~\ref{sec:mmodels}, in signal models where $\manifold$ is
isometric to $\Theta$, this is automatic: we have simply that
\begin{equation*}
\gdist(x_{\widehat{\theta}},x_{\theta^\ast}) =
d_\Theta(\widehat{\theta},\theta^\ast).
\end{equation*}
Such signal models are not nonexistent. Work by Donoho and
Grimes~\cite{DonohoGrimesISOMAP}, for example, has characterized a
variety of articulated image classes for which~(\ref{eq:isometry})
holds or for which $\gdist(x_{\theta_1},x_{\theta_2}) = C_8
d_\Theta(\theta_1, \theta_2)$ for some constant $C_8 > 0$. In
other models it may hold that
$$
C_9 \gdist(x_{\theta_1},x_{\theta_2}) \le d_\Theta(\theta_1,
\theta_2) \le C_{10} \gdist(x_{\theta_1},x_{\theta_2})
$$
for constants $C_9, C_{10} > 0$. Each of these
relationships may be incorporated to the bound (\ref{eq:bound4}).
\section{Conclusions and future work}
\label{sec:concl}
In this paper, we have considered the tasks of signal recovery and
parameter estimation using compressive measurements of a
manifold-modeled signal.
Although these problems differ substantially from the mechanics of
sparsity-based signal recovery, we have seen a number of
similarities that arise due to the low-dimensional geometry of the
each of the concise models.
First, we have seen that a sufficient number of compressive
measurements can guarantee a stable embedding of either type of
signal family, and the requisite number of measurements scales
linearly with the information level of the signal.
Second, we have seen that deterministic instance-optimal bounds in
$\ell_2$ are necessarily weak for both problems.
Third, we have seen that probabilistic instance-optimal bounds in
$\ell_2$ can be derived that give the optimal scaling with respect
to the signal proximity to the concise model and with respect to the
amount of measurement noise.
Thus, our work supports the growing empirical evidence that
manifold-based models can be used with high accuracy in compressive
signal processing.
As discussed in Section~\ref{sec:problem}, there remain several
active topics of research. One matter concerns the problem of
non-differentiable manifolds that arise from certain classes of
articulated image models. Based on preliminary and empirical
work~\cite{mbwPhdThesis,msSmashed}, we believe that a combined
multiscale regularization/measurement process is appropriate for
such problems. However, a suitable theory should be developed to
support this. A second topic of active research concerns fast
algorithms for solving problems such as (\ref{eq:csmrecover}) and
(\ref{eq:parmrecover}). Most successful approaches to date have
combined initial coarse-scale discrete searches with iterative
Newton-like refinements. Due to the problem-specific nuances that
can arise in manifold models, it is unlikely that a single
general-purpose algorithm analogous to $\ell_1$-minimization will
emerge for solving these problems. Nonetheless, advances in these
directions will likely be made by considering existing techniques
for solving (\ref{eq:csmopt}) and (\ref{eq:parmopt}) in the native
space, and perhaps by considering the multiscale measurement
processes described above.
Finally, while we have not considered stochastic models for the
parameter $\theta$ or the noise $\noise$, it would be interesting to
consider these situations as well. A starting point for such
statistical analysis may be the constrained Cram\'{e}r-Rao Bound
formulations~\cite{stoica1998crb,moore2007ccr} in which an unknown
parameter is constrained to live along a low-dimensional manifold.
However, the appropriate approach may once again be
problem-dependent, as the nearest-neighbor estimators
(\ref{eq:parmrecover}), (\ref{eq:parmopt}) we describe can be biased
for nonlinear or non-isometric manifolds.
\section*{Acknowledgements}
The author gratefully acknowledges Rice University, Caltech, and the
University of Michigan, where he resided during portions of this
research. An early version of Theorem~\ref{thm:kappa} appeared in
the author's Ph.D. thesis~\cite{mbwPhdThesis}, under the supervision
of Richard Baraniuk. Thanks to Rich and to the Rice CS research team
for many stimulating discussions.
|
\section{INTRODUCTION}
The discovery of close orbiting extrasolar planets led to extensive
studies of disk planet interactions and the forms of migration that can explain
their location. Early theoretical work established the so-called type I and
type II migration regimes for low mass embedded planets and high mass
gap forming planets
(Goldreich \& Tremaine 1980; Lin \& Papaloizou 1986; Ward 1997), respectively.
Although it is suggested that migration is necessary to account for the observed
distribution of planets (Ida \& Lin 2008),
the problem is that
analytic theories and numerical simulations have shown that migration
timescales of type I are quite short (Tanaka et al. 2002) so that the
planet tends to migrate to its central star before it has time to
become massive enough to open a gap in the disk.
This problem thus becomes a competition between two timescales: type I migration
and core accretion for planet mass growth (Pollack et al. 1996; Hubickyj et al. 2005).
Several mechanisms have been suggested to address this challenging
problem, which include thermal effects of the disk (Jang-Condell
\& Sasselov 2004), radial opacity jump (Menou \& Goodman
2004), magnetic turbulent fluctuations
(Nelson \& Papaloizou 2004) and effects of co-orbital material
(Masset et al. 2006).
Non-isothermal slowing down of type I migration is studied by
Paardekooper \& Mellema (2006),
Baruteau \& Masset (2008), and Kley et al. (2009).
Recently, Li et al. (2009; hereafter Paper I) found that the low mass
planet migration
can have a strong dependence on the disk viscosity. They found that
the type I migration is halted in disks of sufficiently low
viscosity. This is caused by a density feedback effect which results
in a mass redistribution around the planet. The simulations confirm the
existence of a critical mass ($M_{cr} \sim 10 M_{\oplus}$) beyond
which migration is halted in nearly laminar disks. The critical masses
are in good agreement with the analytic model of Rafikov (2002).
This paper is a follow-up study to Paper I. By performing a
series of high resolution, 2-D hydrodynamic simulations, we present a more detailed
analysis on the density feedback effect, and describe the long term
($> 10^4$ orbits) behavior of migration.
The paper is organized as follows. In \S 2 we give a brief description of our simulations.
In \S 3 we discuss the density wave damping mechanism for different
disk viscosities and the consequent long term migration behavior,
including the density feedback and the Rossby vortex instability (RVI).
Possible 3-D effects are discussed in \S 4.
Summary and discussions are given in \S 5. A study on the shock
excitation in inviscid disks is given in the Appendix.
\section{Simulations}
The 2-D hydrodynamic simulation set-up and the numerical methods we
used here are the same as that in Paper I (more details on the code are given in
Li \& Li 2009). We choose an initial surface density profile normalized to
the minimum mass solar nebulae model (Hayashi 1981)
as $\Sigma(r) = 152 f (r/5 AU)^{-3/2} \rm{gm \ cm}^{-2}$, where $f=1$
in this paper. (The migration dependence on $f$ in the low viscosity
limit has been explored in Paper I so we will not vary $f$ in
this study.)
The disk is assumed to be isothermal throughout the simulated region,
having a constant sound speed $c_s$.
The dimensionless disk thickness $c_s/v_{\phi}(r=r_p)$ is set as $0.035$,
where $v_{\phi}$ is the Keplerian velocity at the initial planet location $r_p$.
(Simulations with higher $c_s$ were given in Paper I.)
The dimensionless kinematic viscosity $\nu$ (normalized
by $\Omega^2 r$ at the planet's initial orbital radius) is taken
as a spatial constant and ranges from $0$ to $10^{-6}$. This
corresponds to an effective viscous $\alpha = 1.5\nu/h^2 = 1.2\times 10^3 \nu$.
For most runs, we have chosen the planet mass to be $10 M_{\oplus}$. The planet's Hill
radius is $r_H =0.0215 r_p$, which is $\sim 0.6 h$. A pseudo-3D
softening is used (Li et al. 2005). Fully 2-D disk-self
gravity is included (Li, Buoni, \& Li 2009). The disk is simulated with
$0.4 \leq r \leq 2$. Runs are made typically
using a radial and azimuthal grid of $(n_r\times n_\phi) = 800\times
3200$, though we have used higher resolution to ensure convergence on
some runs. Simulations typically last more than ten thousand orbits so
that we can study the long term behavior of migration.
\section{Results}
Figure \ref{09v_runs} shows the orbital radius evolution of a $10
M_{\oplus}$ protoplanet in a 2-D laminar disk with different disk viscosities.
When the disk viscosity is relatively large ($\nu = 10^{-6}$ or
$\alpha = 1.2\times 10^{-3}$), the migration rate
agrees well with the theoretical results given by Tanaka et al. (2002)
for type I migration. When the disk viscosity is low,
the migration behavior differs markedly from the usual type I
migration (see also Fig. 1 in Paper I). Such slowing down behavior
was explained in terms of the density feedback effect (Ward 1997;
Rafikov 2002) in Paper I. Here, we have extended the evolution to be
about ten times longer
($> 10^4$ orbits) than those in Paper I. But before we discuss the
long term behavior in detail, we present some additional analysis of the
density feedback effect first.
\begin{figure}
\begin{center}
\epsfig{file=09v_runs.eps,height=3in,width=5in,angle=0}
\end{center}
\caption{The orbital radius evolution of a $10 M_{\oplus}$ planet
migrating in disks with
different viscosities $\nu = 0, 10^{-8}, 10^{-7},$ and
$10^{-6}$. The normalized disk sound speed is $c_s=0.035$. The
effective $\alpha$ due to the viscosity is $
\approx 0, 10^{-5}, 10^{-4},$ and $10^{-3}$, respectively.
}
\label{09v_runs}
\end{figure}
\subsection{Density Wave Damping Mechanism for Different Viscosities}
The critical physics issue in deciding the density feedback effect is
the density wave damping mechanism. Where and how the density waves
generated by the protoplanet damp will contribute critically to the
torque on the planet. Furthermore, such damping process will modify
the density distribution around the planet, which directly affects the
torque as well. This effect was partially analyzed in Paper I. In
principle, the density wave can damp both due to disk viscosity (a
viscous process) and by shocks (a nonlinear process). The relative
importance of these processes will naturally depend on the disk
viscosity.
To quantify the damping process, we have evaluated the Reynolds
stress and viscous stress. An effective, azimuthally averaged $\alpha$
based on the Reynolds stress can be defined as:
\begin{equation}
\alpha_{Rey
= \ \left\langle\frac{\Sigma v_r
\delta v_{\phi}}{P}\right\rangle~~,
\end{equation}
where $\langle ... \rangle$ indicates the azimuthal average, $\delta v_{\phi} =
v_{\phi} - \langle v_{\phi}\rangle$, $\Sigma$ and $P$ are disk
surface density and pressure, respectively. This method was previously
discussed in Balbus \& Hawley (1998) and Li et al. (2001).
Similarly, the azimuthally
averaged $\alpha$ based on the viscous stress can be defined as:
\begin{equation}
\alpha_{vis
= \ \left\langle\frac{\nu \Sigma r
\frac{d\Omega}{dr}} {P}\right\rangle~~,
\end{equation}
and this quantity scales as $r^{-3/2}$.
In Figure \ref{reynoldsviscous}, we present both
$\alpha_{Rey}$ and $\alpha_{vis}$ as a
function of disk radius. The results are based on the runs at 400, 1000, 760 and 700 orbits
for $\nu = 10^{-6}, 10^{-7}, 10^{-8}$ and $0$, respectively.
These times are chosen so that the planet is at roughly the same
orbital radius in all the runs.
For $\nu=10^{-6}$, the viscous transport is
much bigger than the Reynolds transport (by more than a factor of 2
around the planet).
For $\nu=10^{-7}$, $ \alpha_{vis} \approx 1.2\times 10^{-4}$, which is
smaller than $\alpha_{Rey}$ around the planet. This means that the
dominant wave damping mechanism around the planet changes from being
viscous damping to being shock-dominated damping when the
disk viscosity changes from $10^{-6}$ to $10^{-7}$. For even smaller
disk viscosity, shocks dominate the wave damping. The peaks of
$\alpha_{Rey}$ are approximately $h (=1.6 r_H)$ away from the
planet, consistent with the excitation of shocks. Note that as disk
viscosity changes, the shock strength and structure will be changed
somewhat. This could account for the changes in $\alpha_{Rey}$ for
$\nu \leq 10^{-7}$.
\begin{figure}
\begin{center}
\epsfig{file=reynoldsViscous2.eps,height=3in,width=5in,angle=0}
\end{center}
\caption{Comparison of the azimuthal averaged Reynold stress
$\alpha_{Rey}$ (solid lines)
and the viscous stress $\alpha_{vis}$ (dashed lines). Note
that $\alpha_{vis}$ roughly scales as
$r^{-3/2}$. For $\nu = 10^{-8}$ and $\nu=0$, $\alpha_{vis}$ is not shown.
For $\nu = 10^{-6}$, the viscous
stress is larger than the Reynolds stress.
For lower values of $\nu = 10^{-7}, 10^{-8}$ and $0$, the viscous stress is
smaller than the Reynolds stress.
The angular momentum transport is shock dominated
when the viscosity $\nu \le 10^{-7}$ or $\alpha \le 10^{-4}$.
}
\label{reynoldsviscous}
\end{figure}
The wave damping by shocks causes the density profiles at the shocks
to be
significantly modified. To confirm this effect further, we analyze the
torque density profiles by examining $dT/dM(r)$ where $T$ is the
torque on the planet by disk material and $M$ is the mass within each
radial ring. We choose three runs with $\nu = 10^{-6}, 10^{-7},
10^{-8}$ and pick the planet radial location at $r_p = 0.945$ to
compare (this corresponds to $t=400, 1000,$ and $760$ orbits for these
runs respectively). From Figure \ref{09v_runs}, their migration trend at this
location is quite different (i.e., the total torque on the planet is
very different). The $dT/dM(r)$ profiles are given in the top panel of Figure
\ref{dTdMcmp09v1v2v3}. For different viscosities, the difference in $dT/dM$ is
not large, within a factor of 2. But the torque amplitude on the planet in
the $\nu=10^{-6}$ case is about a factor of 100 larger than that of the two
cases for $\nu = 10^{-7}$ and $10^{-8}$. This shows that the
difference should be caused by the density variations. The bottom
panel of Figure \ref{dTdMcmp09v1v2v3}
shows the radial disk density profile around the planet. The density
imbalance between the inner Lindblad and outer Lindblad regions are
much stronger for the lower viscosity cases than the case for $\nu = 10^{-6}$.
In the usual picture of wave damping, as the viscosity decreases, the
density waves are expected to propagate farther away (Takeuchi et al. 1996). As a
result, the peak positions of $dT/dM$ are expected to be farther
away from the planet. We did not find such behavior in the
simulation results because the shock dissipation dominates the wave
damping when the viscosity is sufficiently small. This also indicates
that, for these choices of planet mass and disk sound speed, shocks are always
produced.
We want to emphasize that, even though the effective ``viscosity''
caused by shocks is not high (see Figure \ref{reynoldsviscous}), it is
the density imbalance due to the angular momentum transport by the
shocks that causes a big change in the total torque on the planet. For
planet's mass above the critical values described in Rafikov (2002)
and Paper I, the above analysis indicates that there exists a critical
disk viscosity, below which the density wave damping will be dominated
by the shock dissipation and the density feedback effect can slow down
(or halt) the migration. For planet mass less than the critical
values, however, even when the disk viscosity is low enough so that the shock
dissipation is dominant, the density imbalance caused by the shock
dissipation is too weak or taking too long to be able to change the
migration behavior.
\begin{figure}
\begin{center}
\epsfig{file=dTdMcmp09v1v2v3.eps,height=3in,width=5in,angle=0}
\end{center}
\caption{({\it top}) The radial profile of torque density $dT/dM$ for three
different viscosity runs at $t = 400, 1000,$ and $760$ for
$\nu = 10^{-6}, 10^{-7}$, and $10^{-8}$ respectively. The planet is
at $r_p= 0.945$ for all three cases.
({\it bottom}) The azimuthally averaged surface density profiles for
three cases. The imbalance in surface density between the inner and
outer Lindblad regions becomes more pronounced in the lower
viscosity runs.
}
\label{dTdMcmp09v1v2v3}
\end{figure}
\subsection{Long Term Evolution}
The long term evolution ($\sim 10^4$ orbits) for different disk viscosity
(Figure \ref{09v_runs}) is complicated. For three low viscosity cases,
$\nu = 10^{-7}, 10^{-8}$, and $0$, the migration is significantly
slowed down or even reversed, but the detailed behavior is different.
(Note that for $\nu=0$, there is still some low level of numerical
viscosity, which we estimate to be roughly equivalent to $\nu < 10^{-9}$.)
We now discuss each case in detail.
\subsubsection{The $\nu = 10^{-7}$ case}
For $\nu = 10^{-7}$, Figure \ref{09v_runs} shows that the planet
has a steady migration rate.
Figure \ref{vistimescale} shows the comparison between the simulation
and the viscous drift rate calculated using ${\dot r} = 3\nu/2 r_p$.
It looks like that the density feedback effects take the planet
migration into a ``viscous'' limit, where the migration is consistent
with being on the viscous time scale after about 2500 orbits. The
corresponding surface
density profile evolution is shown in Figure \ref{denevol1e-7}. It can
be verified
that the density distribution remains smooth and evolves on the
viscous timescale as well. This implies that the shock damping of the
density waves causes the planet and surrounding disk material to
migrate with approximately the same timescale. This situation is similar to
the previous type II migration study where a gap has formed in the
disk. Upon more detailed analysis, however, the accretion rate throughout
the disk is not quite a constant. This suggests that the steady migration
observed so far could change if we follow it to even longer timescales.
Figure \ref{denevol1e-7} shows that a wide density ``depression'' (not
quite a gap) is forming. Given the wide gap, one might have expected
the excitation of the secondary instability such as the Rossby vortex instability
(RVI), but this instability is suppressed in this case by the disk viscosity.
\begin{figure}
\begin{center}
\epsfig{file=vis-timescale-longer.eps,height=3in,width=5in,angle=0}
\end{center}
\caption{ Solid line is the simulation result for $\nu = 10^{-7}$.
Dashed line is the expected viscous drift rate with $\nu = 10^{-7}$.
The planet migration is consistent with the viscous time scale
after about 2500 orbits.
}
\label{vistimescale}
\end{figure}
\begin{figure}
\begin{center}
\epsfig{file=denevol20-600.eps,height=3in,width=5in,angle=0}
\end{center}
\caption{Surface density (azimuthally averaged)
evolution for the $\nu =10^{-7}$ case. A smooth density
``depression'' forms around
the planet and gradually widens.
}
\label{denevol1e-7}
\end{figure}
\subsubsection{The $\nu = 10^{-8}$ case}
For $\nu = 10^{-8}$, the planet's migration is essentially halted and
gradually going in the reverse direction at late stages. Figure
\ref{den1e-8} shows the density distribution at $t=800$ orbits. The
density ``depression'' is steeper than what was seen in the $\nu =
10^{-7}$ case. For such a low viscosity, the RVI is also excited at a low
level. Figure \ref{den1e-8} shows that
the azimuthal density variation is more pronounced in the low
azimuthal wave number $m$. This is because, during the nonlinear stage
of RVI, vortices will merge (Li et al. 2005), and one is often left
with only large scale variations in disk surface density. (More
detailed discussions on RVI will be given below.)
\begin{figure}
\begin{center}
\epsfig{file=rho40.eps,height=3in,width=5in,angle=0}
\end{center}
\caption{Vortices are excited
for $\nu = 10^{-8}$ due to the Rossby vortex instability.
The 2-D disk surface density, $\Sigma\cdot r^{3/2}$, is shown.
The time is 800 orbits.
}
\label{den1e-8}
\end{figure}
\begin{figure}
\begin{center}
\epsfig{file=newdTdMcmp09v3.eps,height=3in,width=5in,angle=0}
\end{center}
\caption
Variations in $dT/dM$ as RVI is (mildly) influencing the planet
migration during $t \sim 3000 - 3400$ orbits.
The peaks around $\pm h$ correspond to the shock damping.
The drastic variations around $\pm (2-3) h$
are caused by RVI.
}
\label{dTdmcmp09v3}
\end{figure}
Because of the RVI, the behavior of $dT/dM$ becomes more
complicated. This is shown in
Figure \ref{dTdmcmp09v3} where we have plotted the evolution of
$dT/dM$ over a period between $\sim 3000 - 3400$ orbits. This
coincides with a period when the RVI is mildly excited (see Figure \ref{09v_runs}).
The peaks around $\pm h$ in $dT/dM$ are still consistent with the shock damping.
The drastic changes around $\pm (2-3) h$ are due to the
azimuthal asymmetries in surface density caused by RVI, which give
rise to the sign change in $dT/dM$.
We have confirmed that the Lindblad resonance positions for these low
$m$ modes are coincident with the positions where $dT/dM$ change
dramatically around $r-r_p = \pm 2.5 h$, as shown in Figure \ref{dTdmcmp09v3}.
When averaged over a few hundred orbits, however, the changes in
$dT/dM$ cancel out as evident in Figure \ref{dTdMcmp09v1v3}. The
averaged profile, when compared
with that from the $\nu =10^{-6}$ case, shows that RVI causes the
torque contribution to extend to a larger radial extent (the tails
between $\pm (2.5-4) h$), though the peak amplitudes of $dT/dM$ at $\pm
h$ are smaller by about a factor of 4. This implies that the excitation
of RVI has a minor impact on the overall migration in this case.
\begin{figure}
\begin{center}
\epsfig{file=dTdMcmp09v1v3.eps,height=3in,width=5in,angle=0}
\end{center}
\caption{
Comparison of $dT/dM$ profiles for two cases with $\nu =
10^{-6}$ and $\nu = 10^{-8}$ when the planet is at $r_p = 0.939$.
The solid curve is obtained by averaging over hundreds of
orbits.
}
\label{dTdMcmp09v1v3}
\end{figure}
It is not clear why the migration is slowly going outward, nor whether
this trend will continue at much longer times than what was
simulated. This is a regime where both the density feedback effects by shock
dissipation and the influence by mild RVI are playing some roles in planet
migration. Though it seems reasonable to expect that the planet
migration is significantly slowed down when compared to the usual type
I rate, it is difficult to get a definite answer.
\subsubsection{The $\nu = 0$ case}
For $\nu = 0$, the planet migrates in a more complex way, now strongly
influenced by the RVI.
Large amplitude oscillations in the semi-major axis evolution appear and
sometimes exhibit rapid radial drops. Figure \ref{vort} presents several
snapshots of the disk surface density, showing the evolution of RVI.
The vortices exert strong
torques to the planet as they move past the planet. It seems that the
planet's
migration is still inward overall, though it undergoes many oscillations,
reversals, and fast drops (see the black curve in Figure 1).
Several factors could have contributed to
this type of evolution. First, the low disk viscosity makes the
shocks stronger (cf. Figure
\ref{reynoldsviscous}), causing a stronger disk response and faster
disk density evolution. The excitation of RVI is associated with the
inflexion points (which are regions of density depressions) in the
radial profile of potential vorticity (Lovelace et al. 1999; Li et al.
2000). Second, these vortices tend to have slightly different azimuthal speeds
so they will merge (Li et al. 2001; 2005), forming large scale density
structures azimuthally. Third, they are anti-cyclones
with high densities, so they produce their own spiral shocks around
them. Their influence on the surrounding flow and the existence of spiral shocks
lead to an effective angular momentum transport (mostly outward) so
these vortices will gradually migrate inwards (see results also in Li
et al. 2001). This is seen in the current simulations as well. Fourth,
when the vortices migrate away inwardly (on a relatively fast
timescale), the planet migration is subsequently affected because the disk
density profile is significantly changed by these
vortices. Fifth, because the shocks by the planet is
strong, new generations of vortices are produced after the previous
generation has migrated away. All these highly non-linear effects,
unfortunately, make it very difficult to predict the behavior of
planet migration.
\begin{figure}
\begin{center}
\epsfig{file=new-vortex-evol.eps,height=6in,width=6in,angle=0}
\end{center}
\caption{Snapshots of the disk surface density
$\Sigma \cdot r^{3/2}$ at $t = 200$, $600$, $1000$,
$4600$, $5000$, $5400$, $6000$, $8000$, and $9180$ orbits
from the top left panel to
the lower right panel, respectively. The viscosity $\nu =
0$. The planet location is marked by an ``X''. Vortices are
produced as a result of the RVI. These vortices merge,
migrating inward, and being produced anew by the planet.
}
\label{vort}
\end{figure}
One curious observation is the fast radial drop during the planet
migration, as indicated, say, between $t = 6400$ and $8000$ orbits
(cf. Figure \ref{09v_runs}). In Figure \ref{rho-turn398},
we show the density distribution at the time of a rapid drop (7960
orbits). It is interesting to see that the drop is coincident with
this close encounter between the planet and the density blob.
The mass of the dense blob is estimated to be the same order
of the planet, about $3\times 10^{-5}$ or $10 M_{\oplus}$. The $dT/dM$
profiles at the time around the rapid drop are shown in Figure
\ref{dTdMcmp397-399}. We can identify that at the time of 7960 orbits,
a huge negative torque occurs and could contribute to the rapid drop
of the planet.
\begin{figure}
\begin{center}
\epsfig{file=rho-turn398.eps,height=5in,width=5in,angle=0}
\end{center}
\caption{The 2-D disk surface density, $\Sigma\cdot r^{3/2}$, is shown.
A density blob gives rise to a big negative torque at the time of rapid drop
(about 7960 orbits).
}
\label{rho-turn398}
\end{figure}
\begin{figure}
\begin{center}
\epsfig{file=dTdMcmp397-399.eps,height=3in,width=5in,angle=0}
\end{center}
\caption{
Torque density distribution before and after the rapid drop of
the planet ($t \sim 7940 - 7980$ orbits). The large negative
torque appears when the density blob is located lagging
behind planet's orbit.
}
\label{dTdMcmp397-399}
\end{figure}
\section{Possible 3-D Effects}
\label{sec:3-D}
Our results indicate that the disk gas density distribution near the
planet sensitively controls migration. Migration stoppage in low
viscosity disks is a consequence of a systematic mild redistribution
of gas mass near the planet, favoring outward over inward
torques. It does not require complete removal of gas near the planet,
as in the type 2 regime (Li et al 2009). The redistribution is in turn
controlled by shocks. The location and structure of these shocks have an important
influence on the feedback torque on the planet. The nature of the shocks
that occur in 3-D can be quite different from the 2-D case analyzed in
this paper. In a 2-D isothermal disk with pressure, only one type of
wave is excited, a rotationally modified acoustic wave. In a 3-D disk
that is not vertically isothermal and/or has a nonzero vertical buoyancy frequency, this wave is
modified and other types of waves may be excited
(Lubow \& Pringle 1993; Korycansky \& Pringle 1995).
Their damping properties differ from the 2-D case. If the disk is not
vertically isothermal, as suggested by steady-state models of dead
zones (e.g., Terquem 2008), then the
main wave that is excited, the $f$ mode, becomes more confined near the
disk surface as it propagates, through 'wave channelling' (Lubow \& Ogilvie 1998; Bate
et al 2002). The wave becomes more nonlinear as it propagates and undergoes more rapid shock
damping than in the 2-D case. Since the material that gets shocked lies above the disk
midplane, it is not clear how effective the breaking surface waves will be in
affecting migration in comparison to the isothermal case. But, the rate of change of disk
angular momentum produced by waves is determined by the angular momentum flux they
carry. For given disk surface density near a resonance, the $f$ mode carries about the same amount of angular
momentum flux as the 2-D acoustic mode. So if the $f$ mode damps closer to the planet than
the 2-D acoustic mode, then its effects on the migration torque could be more important.
It is possible that the upper layers are successively shocked
from the outside-in towards the midplane
and displaced radially. The process may become less effective, as
the remaining gas becomes less optically thick
and more isothermal. These suggestions are speculative.
Further analysis is required to determine the importance of the $f$ mode
effects on migration.
Modes other than the $f$ mode that are excited in a 3-D disk can damp
rapidly. For a
vertically isothermal disk undergoing adiabatic wave perturbations,
the fraction of the wave
energy that goes into these alternative modes is given by
$1-\sqrt{\gamma(2-\gamma)}$ [see Eq. (B4) in Bate et al 2002].
For $\gamma = 1.4$, this fraction is only about $8\%$. It is
possible that the damping of these waves may produce a feedback torque that is more significant
in strength than $8\%$ of the total feedback torque. The reason is that the wave damping will
likely occur closer to the planet than the 2-D mode investigated in this paper. For example,
vertically propagating gravity waves are produced that damp in the disk atmosphere.
For a $\gamma = 5/3$ gas, the ratio is $25\%$ and the damping effects of these modes are more important. The
damping of these waves occurs well above the disk midplane and
it is not clear how much the feedback torque on the planet is
modified. As discussed above, the disk may be affected from the outside-in, towards the midplane.
A proper 3D analysis of the disk evolution in the low viscosity case
is required.
\section{Summary and Discussion}
\label{sec:diss}
We have carried out 2-D global hydrodynamic simulations to study the
migration of a $10 M_{\oplus}$ protoplanet in a protoplanetary disk.
The disk surface density is taken to have the same value in the minimum
mass solar nebula model, but we have taken the normalized disk sound
speed to be relatively low, $c_s = 0.035$. In Paper I, we have
shown the existence and the concrete values for critical planet masses
(depending on the disk mass and sound speed)
above which the density feedback effects will slow down the type I
migration significantly. Here, we have mainly focused on the long term
behavior of planet migration in such low viscosity disks. We find the
following results:
1) When the disk viscosity is high (e.g., $\nu \geq 10^{-6}$, or $\alpha
\geq 10^{-3}$), the density wave damping is dominated by the disk
viscosity. The migration can be described as the typical type I migration.
2) When the viscosity is relatively low (e.g., $\nu$ is between $\sim
10^{-8}$ and $10^{-6}$, or $\alpha
\sim 10^{-5}$ and $10^{-3}$), the density wave damping is dominated by
shocks. This then modifies the disk surface density profile quite
significantly, which produces a density feedback effect that alters
the planet migration, slowing it down into a viscous time scale or
halting the migration altogether. The new migration timescale, $t
\geq 1/\nu \sim 10^6$ orbits, is considerably longer than the usual
type I migration time. This range of the disk $\alpha$ is
interestingly consistent with the expected values in the ``dead
zone'' of protoplanetary disks where protoplanet cores are believed
to arise. If the cores of protoplanets can manage to grow above the
critical masses (as given in Paper I) without migrating away, then
these cores can spend a long time in the dead zone (essentially the
disk lifetime).
3) When the disk viscosity is even lower (e.g., $\alpha < 10^{-5}$),
the density feedback effect is still present but the RVI starts to
dominate the nonlinear evolution of the disk. The planet migration is
severely affected by the RVI. Large amplitude oscillations appear in the planet
semi-major axis evolution and the rapid drop of the planet occurs sometimes as
RVI-induced density blob experiences close encounters with the
planet. The overall migration seems still inward and becomes
unpredictable. It is not quite clear whether realistic disks will ever
have such low viscosities.
We have only studied the long term migration behavior of a $10
M_{\oplus}$ protoplanet. This mass is above the critical mass limit
discussed in Paper I. For lower planet masses, however, even for low
viscosity disks, the shocks produced by the planet will tend to be
weak, so the density feedback effects discussed in this paper and
Paper I will not be strong enough to slow down the migration
significantly. In this limit, the Type I migration still poses a
serious threat to the survivability of these small mass protoplanets
(say, $< 3 M_{\oplus}$), if no other mechanisms can stop the migration.
The critical masses for stopping planet migration are sensitive to the disk
interior temperature. Dead zones may have higher temperatures than assumed here.
For a steady state disk, the surface density varies inversely with $\nu$.
The higher surface density, due to the lower $\nu$ in a dead zone, gives rise to
a higher optical depth and therefore higher temperature at
the disk midplane (e.g., Terquem 2008). If the disk temperatures reach a
value corresponding to $H/r \ga 0.1$, then the critical masses can become substantially
higher than determined here and the effects of the feedback effect on planet
migration become much less important.
In addition, realistic 3-D simulations are certainly
desirable to address how layered vertical structures
(with both magnetically active and less-active regions)
will affect the wave damping and planet migration.
\acknowledgments
The research at LANL is supported by a Laboratory Directed Research
and Development program. C.Y. thanks the support from National Natural
Science Foundation of China (NSFC, 10703012) and Western Light
Young Scholar Program. S.L. acknowledges support from NASA Origins
grant NNX07AI72G.
|
\section{Introduction}
The discovery of a close relationship between the nature of quantum gravity
and the thermodynamics of black holes has been one of the most important
developments in general relativity in the past decades. Strong motivation
for studying thermodynamics of black holes originates from the fact that
they have a very natural thermodynamic description. For example, black holes
have an entropy and temperature related to their horizon area and surface
gravity, respectively, and also one can investigate their thermal stability.
With the appearance of the anti-de Sitter/conformal field theory
correspondence (AdS/CFT) \cite{ADSCFT1}, such black holes in asymptotically
AdS space become even more interesting since one can gain some significant
relations between the thermodynamical properties of the AdS black holes and
the dual conformal field theory \cite{ADSCFT11,ADSCFT2,ADSCFT3}.
On the other hand, it is a general belief that in four dimensions the
topology of the event horizon of an asymptotically flat stationary black
hole is uniquely determined to be the two-sphere $S^{2}$ \cite{Haw1,Haw2}.
Hawking's theorem requires the integrated Ricci scalar curvature with
respect to the induced metric on the event horizon to be positive \cite{Haw1
. This condition applied to two-dimensional manifolds determines
uniquely the topology. The ``topological censorship theorem'' of
Friedmann, Schleich and Witt is another indication of the
impossibility of non spherical horizons \cite{FSW1,FSW2}. However,
when the asymptotic flatness of spacetime is violated, there is no
fundamental reason to forbid the existence of static or stationary
black holes with nontrivial topologies. It has been shown that for
asymptotically AdS spacetime, in the four-dimensional
Einstein-Maxwell theory, there exist black hole solutions whose
event horizons may have zero or negative constant curvature and
their topologies are no longer the two-sphere $S^{2}$. The
properties of these black holes are quite different from those of
black holes with usual spherical topology horizon, due to the
different topological structures of the event horizons. Besides,
the black hole thermodynamics is drastically affected by the
topology of the event horizon. It was argued that the Hawking-Page
phase transition \cite{Haw3} for the Schwarzschild-AdS black hole
does not occur for locally AdS black holes whose horizons have
vanishing or negative constant curvature, and they are thermally
stable \cite {Birm}. The studies on the topological black holes
have been carried out extensively in many aspects (see e.g.
\cite{Lemos,Cai2,Bril1,Cai33,Cai44}). In this paper we shall
consider topological black branes in the presence of dilaton and
electromagnetic fields in all higher dimensions. The action of the
$(n+1)$-dimensional $(n\geq 3)$ Einstein-Maxwell-dilaton gravity
can be written as
\begin{equation}
I=-\frac{1}{16\pi }\int d^{n+1}x\sqrt{-g}\left( {R}-K(\Phi )-V(\Phi )
\mathcal{L}(\Phi ,F)\right) , \label{Action}
\end{equation}
where ${R}$ is the Ricci scalar\textbf{, }$K(\Phi )=4(\nabla \Phi
)^{2}/(n-1) $ is a kinetic term, $V(\Phi )$ is a potential term for the
dilaton field $\Phi $, and $\mathcal{L}(\Phi ,F)=-e^{-4\alpha \Phi
/(n-1)}F_{\mu \nu }F^{\mu \nu }$ is a coupled Lagrangian between scalar
dilaton and electromagnetic fields. In $\mathcal{L}(\Phi ,F)$, $\alpha $ is
an arbitrary constant governing the strength of the coupling between the
dilaton and the Maxwell field, $F_{\mu \nu }=\partial _{\mu }A_{\nu
}-\partial _{\nu }A_{\mu }$ is the electromagnetic field tensor and $A_{\mu }
$ is the electromagnetic potential. While $\alpha =0$ corresponds to the
usual Einstein-Maxwell-scalar theory, $\alpha =1$ indicates the
dilaton-electromagnetic coupling that appears in the low energy string
action in Einstein's frame.
Some attempts have been made to explore various solutions of
Einstein-Maxwell-dilaton gravity. The dilaton field couples in a nontrivial
way to other fields such as gauge fields and results into interesting
solutions for the background spacetime \cite{CDB1,CDB2}. These scalar
coupled black hole solutions \cite{CDB1,CDB2}, however, are all
asymptotically flat. It was argued that with the exception of a pure
cosmological constant, no dilaton-de Sitter or anti-de Sitter black hole
solution exists with the presence of only one Liouville-type dilaton
potential \cite{MW}. In the presence of one or two Liouville-type
potentials, black hole spacetimes which are neither asymptotically flat nor
(A)dS have been explored by many authors (see e.g. \cite
{CHM,Clem,Sheykhi0,Sheykhi1,Hendi2}). Recently, the \textquotedblleft
cosmological constant term\textquotedblright\ in the dilaton gravity has
been found by Gao and Zhang \cite{Gao}. With an appropriate combination of
three Liouville-type dilaton potentials, they obtained the static dilaton
black hole solutions which are asymptotically (A)dS in four and higher
dimensions. In such a scenario AdS spacetime constitutes the vacuum state
and the black hole solution in such a spacetime becomes an important area to
study \cite{ADSCFT1}. For an arbitrary value of $\alpha $ in AdS spaces the
form of the dilaton potential ${V}({\Phi })$ in $n+1$ dimensions is chosen
as \cite{Gao}
\begin{eqnarray}
{V}({\Phi }) &=&\frac{2\Lambda }{n(n-2+\alpha ^{2})^{2}}\Bigg\{-\alpha ^{2
\left[ (n+1)^{2}-(n+1)\alpha ^{2}-6(n+1)+\alpha ^{2}+9\right] e^{-4(n-2)
\Phi /}[(n-1)\alpha ]} \nonumber \\
&&+(n-2)^{2}(n-\alpha ^{2})e^{4\alpha {\Phi /}(n-1)}+4\alpha
^{2}(n-1)(n-2)e^{-2{\Phi }(n-2-\alpha ^{2})/[(n-1)\alpha ]}\Bigg \},
\label{V(Phi)}
\end{eqnarray}
where $\Lambda $ is the cosmological constant. The motivations for
studying such dilaton black holes with nonvanishing cosmological
constant originate from supergravity theory. Gauged supergravity
theories in various dimensions are obtained with negative
cosmological constant in a supersymmetric theory. In addition, it
has been shown that one may consider a Big Bang model of the
Universe in the presence of dilaton field and presented
Liouville-type potential which can mimic the matter (including
dark matter) and dark energy. This model predict age of the
Universe, transition redshift, Big Bang nucleosynthesis and
evolution of dark energy agree with current observations
\cite{GZH}. Also, this type of potential can be obtained when a
higher dimensional theory is compactified to four dimensions,
including various super gravity models \cite{RaduTch} (see also
\cite{Giddings} for a recent discussion of these aspects). In
particular, for special values of coupling constant, $\alpha$,
this potential reduce to the supersymmetry potential of Gates and
Zwiebach in string theory \cite{RaduTch}.
For later convenience we redefine $\Lambda =-n(n-1)/2l^{2}$, where $l$ is
the AdS radius of spacetime. It is clear the cosmological constant is
coupled to the dilaton in a very nontrivial way. In the absence of the
dilaton field action (\ref{Action}) reduces to the action of
Einstein-Maxwell gravity with cosmological constant. Considering this type
of dilaton potential, one can extract successfully the AdS solutions of
Einstein--Maxwell-dilaton gravity \cite{Sheykhi2,SheDehHen2010}.
The rest of this paper is outlined as follows. In the next section, we
consider the field equations of Einstein-Maxwell-dilaton gravity and present
the $(n+1)$-dimensional topological AdS black brane solutions and investigate
their properties. In section \ref{Therm}, we obtain the conserved and
thermodynamic quantities of the solutions and verify the validity of the
first law of black brane thermodynamics. We perform a stability analysis in
the canonical ensemble and disclose the effect of the dilaton field on the
thermal stability of the solutions in section \ref{Stab}. Conclusions are
drawn in the last section.
\section{Field equations and solutions\label{Field}}
Varying action (\ref{Action}) with respect to the metric tensor $g_{\mu \nu
} $, the dilaton field $\Phi $ and the electromagnetic potential $A_{\mu }$,
the equations of motion are obtained as
\begin{equation}
{R}_{\mu \nu }=\frac{4}{n-1}\left( \partial _{\mu }\Phi \partial _{\nu }\Phi
+\frac{1}{4}g_{\mu \nu }V(\Phi )\right) +2e^{-4\alpha \Phi /(n-1)}F_{\mu
\eta }F_{\nu }^{\text{ }\eta }-\frac{g_{\mu \nu }}{(n-1)}\mathcal{L}(\Phi
,F), \label{FE1}
\end{equation}
\begin{equation}
\nabla ^{2}\Phi =\frac{n-1}{8}\frac{\partial V}{\partial \Phi }-\frac{\alpha
}{2}\mathcal{L}(\Phi ,F), \label{FE2}
\end{equation}
\begin{equation}
\nabla _{\mu }\left( e^{-{4\alpha \Phi }/({n-1})}F^{\mu \nu }\right) =0.
\label{FE3}
\end{equation}
Here we want to obtain the $(n+1)$-dimensional static solutions of Eqs. (\ref
{FE1}), (\ref{FE2}) and (\ref{FE3}). We assume that the metric has the
following form
\begin{equation}
d{s}^{2}=-N^{2}(\rho )f^{2}(\rho )dt^{2}+\frac{d\rho ^{2}}{f^{2}(\rho )
+\rho ^{2}{R^{2}(\rho )}d\Omega _{n-1}^{2}, \label{metric}
\end{equation}
where
\[
d\Omega _{n-1}^{2}=\left\{
\begin{array}{cc}
d\theta _{1}^{2}+\sinh ^{2}\theta _{1}d\theta _{2}^{2}+\sinh ^{2}\theta
_{1}\sum\limits_{i=3}^{n-1}\prod\limits_{j=2}^{i-1}\sin ^{2}\theta
_{j}d\theta _{i}^{2} & k=-1 \\
\sum\limits_{i=1}^{n-1}d\phi _{i}^{2} & k=0
\end{array}
\right.
\]
represents the line element of an $(n-1)$-dimensional hypersurface with
constant curvature $(n-1)(n-2)k$. It is notable
that positive curvature horizon ($k=1$) has been investigated in \cite
{SheDehHen2010}. Here $N(\rho )$, $f(\rho )$ and $R(\rho )$ are functions of
$\rho $ which should be determined. First of all, the Maxwell equations (\ref
{FE3}) can be integrated immediately, where all the components of ${F}_{\mu
\nu }$ are zero except ${F}_{t\rho }
\begin{equation}
{F}_{t\rho }=N(\rho )\frac{qe^{4\alpha {\Phi /(n-1)}}}{\left( \rho R\right)
^{n-1}}, \label{Ftr}
\end{equation}
where $q$, an integration constant, is the charge parameter of the black
brane. Our aim here is to construct exact, $(n+1)$-dimensional topological
AdS black brane solutions of Eqs. (\ref{FE1})-(\ref{FE3}), with the dilaton
potential (\ref{V(Phi)}) for an arbitrary dilaton coupling parameter $\alpha
$ and investigate their properties. Using metric (\ref{metric}) and the
Maxwell field (\ref{Ftr}), one can show that the system of equations (\ref
{FE1})-(\ref{FE2}) have solutions of the form
\begin{eqnarray}
N^{2}(\rho ) &=&\Upsilon ^{-\gamma (n-3)}, \label{Nrho} \\
f^{2}(\rho ) &=&\frac{\rho ^{2}}{l^{2}}\Upsilon ^{(n-2)\gamma }+\left[
k-\left( \frac{c}{\rho }\right) ^{n-2}\right] \Upsilon ^{1-\gamma },
\label{frho} \\
{\Phi }(\rho ) &=&\frac{n-1}{4}\sqrt{\gamma (2+2\gamma -n\gamma )}\ln
\Upsilon , \label{Phirho} \\
R^{2}(\rho ) &=&\Upsilon ^{\gamma }, \label{Rrho} \\
\Upsilon &=&1-\left( \frac{b}{\rho }\right) ^{n-2}. \nonumber
\end{eqnarray}
Here $c$ and $b$ are integration constants and the constant $\gamma $ is
\begin{equation}
\gamma =\frac{2\alpha ^{2}}{(n-2)(n-2+\alpha ^{2})}. \label{gamma}
\end{equation}
The charge parameter $q$ is related to $b$ and $c$ by
\begin{equation}
q^{2}=\frac{(n-1)(n-2)^{2}}{2(n-2+\alpha ^{2})}c^{n-2}b^{n-2}. \label{q}
\end{equation}
For $\alpha \neq 0$ the solutions are not real for $0<\rho <b$ and therefore
we should exclude this region from the spacetime. For this purpose we
introduce the new radial coordinate $r$ as
\begin{equation}
r^{2}=\rho ^{2}-b^{2}\Rightarrow d\rho ^{2}=\frac{r^{2}}{r^{2}+b^{2}}dr^{2}.
\label{Transformation}
\end{equation}
With this new coordinate, the above metric becomes
\begin{equation}
d{s}^{2}=-N^{2}(r)f^{2}(r)dt^{2}+\frac{r^{2}dr^{2}}{(r^{2}+b^{2})f^{2}(r)
+(r^{2}+b^{2}){R^{2}(r)}d\Omega _{n-1}^{2}, \label{Metric2}
\end{equation}
where the coordinates $r$ assumes the values $0\leq r<\infty $, and
N^{2}(r) $, $f^{2}(r)$, $\Phi (r)$ and $R^{2}(r)$ are now given as
\begin{eqnarray} \label{gr}
N^{2}(r) &=&\Gamma ^{-\gamma (n-3)}, \label{Nr} \\
f^{2}(r) &=&\frac{r^{2}+b^{2}}{l^{2}}\Gamma ^{(n-2)\gamma }+\left[ k-\left(
\frac{c}{\sqrt{r^{2}+b^{2}}}\right) ^{n-2}\right] \Gamma ^{1-\gamma },
\label{fr} \\
{\Phi }(r) &=&\frac{n-1}{4}\sqrt{\gamma (2+2\gamma -n\gamma )}\ln \Gamma ,
\label{Phir} \\
R^{2}(r) &=&\Gamma ^{\gamma }, \label{Rr} \\
\Gamma &=&1-\left( \frac{b}{\sqrt{r^{2}+b^{2}}}\right) ^{n-2}. \nonumber
\end{eqnarray}
\subsection{Properties of the solutions:}
It is notable to mention that these solutions are valid for all values of
\alpha $. When ($\alpha =0=\gamma $), these solutions describe the $(n+1)
-dimensional asymptotically AdS Reissner-Nordstrom black branes. For $b=0$
(\Gamma =1)$, the charge parameter $q$ and the scalar field ${\Phi }(r)$
vanish and our solutions reduce to the solutions of Einstein gravity in the
presence of cosmological constant.
Here we should discuss singularity(ies). After some algebraic
manipulation, one can show that the Kretschmann and the Ricci scalars in
(n+1)$-dimensions are finite for $r\neq 0$, but in the vicinity of $r=0$, we
have
\begin{eqnarray*}
R_{\mu \nu \lambda \kappa }R^{\mu \nu \lambda \kappa } &\propto &r^{-4(n-2)
\left[ 1+\left( \alpha ^{2}+n-2\right) ^{-1}\right] }, \\
R &\propto &r^{-2(n-2)\left[ 1+\left( \alpha ^{2}+n-2\right) ^{-1}\right] },
\end{eqnarray*}
and thus they diverge at $r=0$ ($\rho =b$). Thus, $r=0$ is a curvature
singularity. It is worthwhile to note that the scalar field $\Phi (r)$ and
the electromagnetic field $F_{t\rho }$ become zero as $r\rightarrow \infty
(\rho \rightarrow \infty )$. One should note that for nonzero $\alpha$ the
singularity may be null, while it is timelike for $\alpha =0$ (see Figs. \ref
{Figf21} and \ref{Figf221}). Also, figure \ref{Figf221} shows that depending
on the metric parameters, these real solutions may be interpreted as black
brane solutions with inner and outer horizons, an extreme black brane or naked
singularity.
\begin{figure}[tbp]
\epsfxsize=8cm \centerline{\epsffile{f2alpha0.eps}}
\caption{$f^{2}(\protect\rho )$ versus $\protect\rho $ for $b=1$, $c=1.5$,
l=0.5$, $n=5$ and $\protect\alpha =0$ (timelike singularity). $k=0$ (bold
line), and $k=-1$ (dashed line).}
\label{Figf21}
\end{figure}
\vspace{0.2cm}
\begin{figure}[tbp]
$
\begin{array}{cc}
\epsfxsize=8cm \epsffile{f21k0.eps} & \epsfxsize=8cm \epsffile{f21kneg1.eps}
\end{array}
\caption{$f^{2}(r)$ versus $r$ for $b=1$, $c=2.5$, $l=0.5$ and $n=5$.
\protect\alpha =0.1$ (solid line: naked singularity), $\protect\alpha
\protect\alpha _{ext}$ (bold line: extreme black brane), $\protect\alpha =0.8$
(dashed line: null singularity with two horizons), and $\protect\alpha =1.2$
(dotted line: null singularity with one horizon), where $\protect\alpha
_{ext}=0.59,0.67$ for $k=-1,0$, respectively.}
\label{Figf221}
\end{figure}
\section{Thermodynamics of AdS dilaton black brane\label{Therm}}
In this section we intend to study thermodynamics of topological dilaton
black branes in the background of AdS spaces. First of all we focus on
entropy. The entropy of the dilaton black hole typically satisfies the
so-called area law of the entropy which states that the entropy of the black
hole is one-quarter of the event horizon area \cite{Beck}. This near
universal law applies to almost all kinds of black objects, including
dilaton black holes, in Einstein gravity \cite{hunt}. It is a matter of
calculation to show that the entropy of the dilaton black brane per unit volume is
\begin{equation}
{S}=\frac{b^{n-1}\Gamma _{+}^{\gamma (n-1)/2}}{4\left( 1-\Gamma
_{+}\right) ^{(n-1)/(n-2)}}, \label{entropy}
\end{equation}
where $\Gamma _{+}=\Gamma (r=r_{+})$. It is notable to mention that in
contrast with the higher derivative gravities that may lead to negative
entropy \cite{negativeS}, the presented entropy is positive definite (
0\leq \Gamma _{+}\leq 1$). The Hawking temperature of the dilaton black brane
on the outer horizon $r_{+}$, may be obtained through the use of the
definition of surface gravity,
\begin{equation}
T_{+}=\frac{1}{2\pi }\sqrt{-\frac{1}{2}\left( \nabla _{\mu }\chi _{\nu
}\right) \left( \nabla ^{\mu }\chi ^{\nu }\right) }=\sqrt{r_{+}^{2}+b^{2}
\left. \frac{\left( N^{2}f^{2}\right) ^{^{\prime }}}{4\pi Nr}\right|
_{r=r_{+}}, \label{T}
\end{equation}
where $\chi =\partial _{t}$ is the killing vector and a prime stands for the
derivative with respect to $r$. Finding the radius of outer horizon in terms
of the parameters of the metric function is not possible analytically, and
therefore we obtain the constant $c$ in terms of $b$, $\alpha $ and $r_{h}$
by solving $f(r_{h})=0$, where $r_{h}$ is the radius of inner or outer
horizon of the black brane. Substituting $c$ into Eq. (\ref{T}), one obtains
\begin{equation}
{T}_{h}=\frac{b(n-2)\Gamma _{h}^{1-\gamma (n-1)/2}}{4\pi \left( 1-\Gamma
_{h}\right) ^{1/(n-2)}}\Bigg\{\frac{\left[ \gamma (n-1)-1\right] }
l^{2}\Gamma _{h}^{2-\gamma (n-1)}}-\frac{(n-1)\left[ \gamma (n-2)-2\right] }
(n-2)l^{2}\Gamma _{h}^{1-\gamma (n-1)}}+\frac{k(1-\Gamma _{h})^{2/(n-2)}}
b^{2}}\Bigg\}, \label{Tem}
\end{equation}
where $\Gamma _{h}=\Gamma (r=r_{h})$.
\begin{figure}[tbp]
$
\begin{array}{cc}
\epsfxsize=8cm \epsffile{fig1Tk0.eps} & \epsfxsize=8cm
\epsffile{fig1Tkneg1.eps}
\end{array}
\caption{T versus $r_{+}$ for $b=0.2$, $l=0.5$ and $\protect\alpha =0.1,$
n=4$ (solid line), $n=5$ (bold line), and $n=6$ (dashed line).}
\label{Fig1}
\end{figure}
\vspace{0.2cm}
\begin{figure}[tbp]
$
\begin{array}{cc}
\epsfxsize=8cm \epsffile{fig2Tk0.eps} & \epsfxsize=8cm
\epsffile{fig2Tkneg1.eps}
\end{array}
\caption{T versus $r_{+}$ for $b=0.2$, $l=0.5$ and $\protect\alpha =2,$ $n=4$
(solid line), $n=5$ (bold line), and $n=6$ (dashed line).}
\label{Fig11}
\end{figure}
\vspace{0.2cm}
\begin{figure}[tbp]
$
\begin{array}{cc}
\epsfxsize=8cm \epsffile{fig3Tk0.eps} & \epsfxsize=8cm
\epsffile{fig3Tkneg1.eps}
\end{array}
\caption{T versus $r_{+}$ for $b=0.2$, $l=0.5$ and $\protect\alpha =5,$ $n=4$
(solid line), $n=5$ (bold line), and $n=6$ (dashed line).}
\label{Fig111}
\end{figure}
\vspace{0.2cm}
\begin{figure}[tbp]
$
\begin{array}{cc}
\epsfxsize=8cm \epsffile{fig4Tk0.eps} & \epsfxsize=8cm
\epsffile{fig4Tkneg1.eps}
\end{array}
\caption{T versus $r_{+}$ for $b=0.2$, $l=0.5$ and $n=4$, $\protect\alpha
=0.5$ (solid line), $\protect\alpha =1.5$ (bold line), and $\protect\alpha
=2.5$ (dashed line).}
\label{Fig1111}
\end{figure}
The equation $T_{h}=0$ has
one real root for $k=0$:
\begin{equation}
r_{\mathrm{ext}}=b\left\{ \left( \frac{(n-1)\left[ 2-\gamma (n-2)\right] }{n
\right) ^{2/(n-2)}-1\right\} ^{1/2}, \label{rext}
\end{equation}
while it may have two real roots ($r_{\mathrm{1ext}}$\ and $r_{\mathrm{2ext}}
$) for $k=-1$. These roots are the radius of the extreme black branes. We are
interested in the thermodynamics of event horizon, $r_{+}$, of the black
branes and therefore we consider $T_{+}\geq 0$. The negative values of $T_{h}
\ associated to the temperature of inner horizon. The radius of event
horizon $r_{+}\geq r_{\mathrm{ext}}$\ for $k=0$, and $r_{+}\leq r_{\mathrm
1ext}}$\ or $r_{+}\geq r_{2\mathrm{ext}}$\ for $k=-1$. These facts can be
seen in Figs. \ref{Fig1} - \ref{Fig1111}, which show the temperature versus
r_{+}$\ in various dimensions. In order to be more clear, we discuss these
figures for $k=0$\ and $k=-1$, separately.
\textbf{$k=0$}:
As one can see in Figs. \ref{Fig1} - \ref{Fig1111}, there exist extreme
black brane with radius $r_{\mathrm{ext}}$\ provided $\alpha <\alpha _
\mathrm{ext}}$, where
\begin{equation}
\alpha _{\mathrm{ext}}^{2}=\frac{n-2}{n}\left( n-2+2(n-1)\left[ \left( 1
\frac{r_{\mathrm{ext}}^{2}}{b^{2}}\right) ^{1-n/2}-1\right] \right) .
\label{alphaCrit}
\end{equation}
That is the dilaton field removes the existence of extreme black branes.
\textbf{$k=-1$}:
For black branes with hyperbolic horizon and medium values of
$\alpha $, numerical analysis shows that the equation $T_{h}=0$ \
has two real roots and therefore one can have both small
(r$_{+}\leq r_{\mathrm{1ext}}$) and large (r$_{+}\geq
r_{2\mathrm{ext}}$) black branes. This can be seen on Figs.
\ref{Fig1} - \ref{Fig1111}. Concerning the metric function $f(r)$
and temperature $T$, Figs. \ref{Figf221} - \ref{Fig1111}, one can
find that for small values of $\alpha$ they behave like
charged-AdS black branes while for large values of coupling
constant $\alpha$, they are approximately Schwarzschild-AdS black
branes.
Inserting solutions (\ref{frho})-(\ref{Rrho}) in Eq. (\ref{Ftr}), with
considering the new coordinate (\ref{Transformation}), the electromagnetic
field can be simplified as
\begin{equation}
F_{tr}=\frac{q}{\left( r^{2}+b^{2}\right) ^{\left( n-1\right) /2}}.
\label{Ftrho2}
\end{equation}
As one can see from Eq. (\ref{Ftrho2}), in the background of AdS universe,
the dilaton field does not exert any direct influence on the matter field
F_{tr}$, however, the dilaton field modifies the geometry of the spacetime
as it participates in the field equations. This is in contrast to the
solutions presented in \cite{CHM,Clem,Sheykhi0,Sheykhi1}. The solutions of
Ref. \cite{CHM,Clem,Sheykhi0,Sheykhi1} are neither asymptotically flat nor
(A)dS and the gauge field crucially depends on the scalar dilaton field.
The electric charge of the black brane per unit volume, $Q$, can be found by calculating the
flux of the electromagnetic field at infinity (Gauss theorem), obtaining
\begin{equation}
Q=\frac{1}{4\pi }\int_{\rho \rightarrow \infty }d^{n-1}x\sqrt{-{g}}{F
_{t\rho }=\frac{\Omega _{n-1}}{4\pi }q. \label{Q}
\end{equation}
Let us return to Eq. (\ref{Ftrho2}). The gauge potential $A_{t}$
corresponding to the electromagnetic field (\ref{Ftrho2}) can be easily
obtained as
\[
A_{t}=-\frac{q}{(n-2)\left( r^{2}+b^{2}\right) ^{\left( n-2\right) /2}}.
\]
The electric potential $U$, measured at infinity with respect to the
horizon, is defined by \cite{Cal}
\begin{equation}
U=A_{\mu }\chi ^{\mu }\left| _{r\rightarrow \infty }-A_{\mu }\chi ^{\mu
}\right| _{r=r_{+}}, \label{Pot}
\end{equation}
where $\chi =\partial _{t}$ is the null generator of the horizon. Therefore,
the electric potential may be obtained as
\begin{equation}
U=\frac{q}{(n-2)\left( r_{+}^{2}+b^{2}\right) ^{\left( n-2\right) /2}}.
\label{pot}
\end{equation}
The quasilocal mass of the dilaton AdS black hole can be calculated through
the use of the subtraction method of modified Brown and York (BY) \cite{BY}. Such a
procedure causes the resulting physical quantities to depend on the choice
of reference background. In order to use the BY method the metric should
have the form
\begin{equation}
ds^{2}=-W({\mathcal{R}})dt^{2}+\frac{d\mathcal{R}^{2}}{V(\mathcal{R})}
\mathcal{R}^{2}d\Omega ^{2}. \label{Mets}
\end{equation}
Thus, we should write the metric (\ref{metric}) in the above form. To do
this, we perform the following transformation \cite{DB}:
\[
\mathcal{R}=\rho \Upsilon ^{\gamma /2}.
\]
It is a matter of calculations to show that the metric (\ref{metric}) may be
written as (\ref{Mets}) with the following $W$ and $V$:
\begin{eqnarray*}
W(\mathcal{R}) &=&N^{2}(\rho (\mathcal{R}))f^{2}(\rho (\mathcal{R})), \\
V(\mathcal{R)} &=&f^{2}(\rho (\mathcal{R}))\left( \frac{d\mathcal{R}}{d\rho
\right) ^{2}=\Upsilon ^{(\gamma -2)}\left[ 1+\frac{1}{2}\left( \gamma
(n-2)-2\right) \left( 1-\Upsilon \right) \right] ^{2}f^{2}(\rho (\mathcal{R
)).
\end{eqnarray*}
The background metric is chosen to be the metric (\ref{Mets}) with
\begin{equation}
W_{0}(\mathcal{R})=V_{0}(\mathcal{R})=f_{0}^{2}(\rho (\mathcal{R}))=k+\frac
\rho ^{2}}{l^{2}}+\left\{
\begin{array}{cc}
-\frac{2b\rho \alpha ^{2}}{(1+\alpha ^{2})l^{2}}+\frac{b^{2}\alpha ^{4}}
(1+\alpha ^{2})^{2}l^{2}} & \text{for \ }n=3 \\
-\frac{b^{2}\alpha ^{2}}{(2+\alpha ^{2})l^{2}} & \text{for \ }n=4 \\
0 & \text{for \ }n\geqslant 5
\end{array}
\right. \label{W0}
\end{equation}
As one can see from the above equation, the solutions for $n=3$
and $n=4$ have not ''exact'' asymptotic AdS behavior. Because of
this point, we cannot use the AdS/CFT correspondence to compute
the mass. Indeed, for $n\geqslant 5 $ the metric is exactly
asymptotically AdS, while for $n=3,4$ it is \textit{approximately}
asymptotically AdS. This is due to the fact that if one computes
the Ricci scalar then it is not equal to $-n(n+1)/l^{2}$. It is
well-known that the Ricci scalar for AdS spacetime should have
this value (see e.g. \cite{Weinberg}). Also, the metrics with
$f_{0}^{2}(\rho )$ given by Eq. (\ref{W0}) for $n=3$ and $n=4$ do
not satisfy the Einstein equation with the cosmological constant,
while an AdS spacetime should satisfy the
Einstein equation with cosmological constant. On the other side, at large
\rho $, the metric behaves as $\rho ^{2}$ in all dimensions and therefore we
used the word \textit{''approximately''} asymptotically AdS.
To compute the conserved mass of the spacetime, we choose a timelike Killing
vector field $\xi $ on the boundary surface $\mathcal{B}$ of the spacetime
\ref{Mets}). Then the quasilocal conserved mass can be written as
\begin{equation}
\mathcal{M}=\frac{1}{8\pi }\int_{\mathcal{B}}d^{2}\varphi \sqrt{\sigma
\left\{ \left( K_{ab}-Kh_{ab}\right) -\left(
K_{ab}^{0}-K^{0}h_{ab}^{0}\right) \right\} n^{a}\xi ^{b},
\end{equation}
where $\sigma $ is the determinant of the metric of the boundary $\mathcal{B}
$, $K_{ab}^{0}$ is the extrinsic curvature of the background metric and
n^{a}$ is the timelike unit normal vector to the boundary $\mathcal{B}$. In
the context of counterterm method, the limit in which the boundary $\mathcal
B}$ becomes infinite ($\mathcal{B}_{\infty }$) is taken, and the counterterm
prescription ensures that the action and conserved charges are finite.
Although the explicit function $f(\rho (\mathcal{R}))$ cannot be obtained,
but at large $\mathcal{R}$ this can be done. Thus, one can calculate the
mass per unit volume through the use of the above modified Brown and York formalism as
\begin{equation}
{M}=\frac{n-1}{16\pi }\left[ c^{n-2}+k\left( \frac{n-2-\alpha
^{2}}{n-2+\alpha ^{2}}\right) b^{n-2}\right] . \label{Mass}
\end{equation}
In the absence of a non-trivial dilaton field ($\alpha =0$), this expression
for the mass reduces to the mass of the $(n+1)$-dimensional asymptotically
AdS black brane.
Finally, we check the first law of thermodynamics for the black brane. In
order to do this, we obtain the mass $M$ as a function of extensive
quantities $S$ and $Q$. Using the expression for the charge, the mass and
the entropy given in Eqs. (\ref{Q}), (\ref{Mass}) and (\ref{entropy}), we
can obtain a Smarr-type formula per unit volume as
\begin{equation}
M(S,Q)=\frac{(n-1)}{16\pi }\left[ \frac{32\pi ^{2}(n-2+\alpha
^{2})Q^{2}b^{2-n}}{(n-1)(n-2)^{2}}+k\left( \frac{n-2-\alpha ^{2}}{n-2+\alpha
^{2}}\right) b^{n-2}\right] , \label{Msmarr}
\end{equation}
where $b=b(Q,S)$. One may then regard the parameters $S$ and $Q$ as a
complete set of extensive parameters for the mass $M(S,Q)$ and define the
intensive parameters conjugate to $S$ and $Q$. These quantities are the
temperature and the electric potential
\begin{eqnarray}
T &=&\left( \frac{\partial {M}}{\partial {S}}\right) _{Q}=\frac{\left( \frac
\partial {M}}{\partial {b}}\right) _{Q}\left( \frac{\partial {b}}{\partial {
}_{+}}\right) _{Q}}{\left( \frac{\partial {S}}{\partial {r}_{+}}\right)
_{Q}+\left( \frac{\partial {S}}{\partial {b}}\right) _{Q}\left( \frac
\partial {b}}{\partial {r}_{+}}\right) _{Q}}, \label{inte1} \\
U &=&\left( \frac{\partial {M}}{\partial {Q}}\right) _{S}=\left( \frac
\partial {M}}{\partial {Q}}\right) _{S}+\left( \frac{\partial {M}}{\partial
b}}\right) _{S}\left( \frac{\partial {Q}}{\partial {b}}\right) _{S},
\label{inte2}
\end{eqnarray}
where
\begin{eqnarray}
&&\left( \frac{\partial {b}}{\partial {r}_{+}}\right) _{Q}=-\frac{\left(
\frac{\partial {Z}}{\partial {r}_{+}}\right) _{Q}}{\left( \frac{\partial {Z
}{\partial {b}}\right) _{Q}}, \\
&&Z=\left[ k-\frac{32\pi ^{2}(n-2+\alpha ^{2})Q^{2}}{(n-1)(n-2)^{2}b^{2n-4}
\left( \frac{b}{r_{+}}\right) ^{n-2}\right] \left[ 1-\left( \frac{b}{r_{+}
\right) ^{n-2}\right] ^{1-\gamma \left( n-2\right) } \nonumber \\
&&-\frac{r_{+}^{2}}{l^2}\left[ 1-\left( \frac{b}{r_{+}}\right) ^{n-2}\right]
^{\gamma }.
\end{eqnarray}
Straightforward calculations show that the intensive quantities calculated
by Eqs. (\ref{inte1}) and (\ref{inte2}) coincide with Eqs. (\ref{Tem}) and
\ref{pot}). Thus, these thermodynamics quantities satisfy the first law of
black brane thermodynamics,
\begin{equation}
dM=TdS+Ud{Q}.
\end{equation}
\vspace{0.2cm}
\begin{figure}[tbp]
$
\begin{array}{cc}
\epsfxsize=8cm \epsffile{fig1HCk0.eps} & \epsfxsize=8cm
\epsffile{fig1HCkneg1.eps}
\end{array}
\caption{$10^{-6}(\partial ^{2}M/\partial S^{2})_{Q}$ (solid line) and $T$
(dashed line) versus $\protect\alpha $ for $b=0.2$, $n=4$, $l=1$ and
r_{+}=0.001$.}
\label{Fig3}
\end{figure}
\vspace{0.2cm}
\begin{figure}[tbp]
$
\begin{array}{cc}
\epsfxsize=8cm \epsffile{fig2HCk0.eps} & \epsfxsize=8cm
\epsffile{fig2HCkneg1.eps}
\end{array}
\caption{$10^{-5}(\partial ^{2}M/\partial S^{2})_{Q}$ (solid line) and $T$
(dashed line) versus $\protect\alpha $ for $b=0.2$, $n=4$, $l=1$ and
r_{+}=0.01$.}
\label{Fig4}
\end{figure}
\vspace{0.2cm}
\begin{figure}[tbp]
$
\begin{array}{cc}
\epsfxsize=8cm \epsffile{fig3HCk0.eps} & \epsfxsize=8cm
\epsffile{fig3HCkneg1.eps}
\end{array}
\caption{$10^{-2}(\partial ^{2}M/\partial S^{2})_{Q}$ (solid line) and $T$
(dashed line) versus $\protect\alpha $ for $b=0.2$, $n=4$, $l=1$ and
r_{+}=0.1$.}
\label{Fig5}
\end{figure}
\vspace{0.2cm}
\begin{figure}[tbp]
$
\begin{array}{cc}
\epsfxsize=8cm \epsffile{fig5HCk0.eps} & \epsfxsize=8cm
\epsffile{fig5HCkneg1.eps}
\end{array}
\caption{$10^{2}(\partial ^{2}M/\partial S^{2})_{Q}$ (solid line) and $T$
(dashed line) versus $\protect\alpha $ for $b=0.2$, $n=4$, $l=1$ and
r_{+}=5 $.}
\label{Fig6}
\end{figure}
\section{Stability in the canonical ensemble\label{Stab}}
Finally, we study the thermal stability of the solutions in the canonical
ensemble. In particular, we will see that the scalar dilaton field makes the
solution unstable. The stability of a thermodynamic system with respect to
small variations of the thermodynamic coordinates is usually performed by
analyzing the behavior of the entropy $S(M,Q)$ around the equilibrium. The
local stability in any ensemble requires that $S(M,Q)$ be a convex function
of the extensive variables or its Legendre transformation must be a concave
function of the intensive variables. The stability can also be studied by
the behavior of the energy $M(S,Q)$ which should be a convex function of its
extensive variable. Thus, the local stability can in principle be carried
out by finding the determinant of the Hessian matrix of $M(S,Q)$ with
respect to its extensive variables $X_{i}$, $\mathbf{H}_{X_{i}X_{j}}^{M}=
\partial ^{2}M/\partial X_{i}\partial X_{j}]$ \cite{Cal,Gub}. In our case
the mass $M$ is a function of entropy and charge. The number of
thermodynamic variables depends on the ensemble that is used. In the
canonical ensemble, the charge is a fixed parameter and therefore the
positivity of the $(\partial ^{2}M/\partial S^{2})_{Q}$ is sufficient to
ensure local stability. In Figs. \ref{Fig3} - \ref{Fig6} we show the
behavior of the $(\partial ^{2}M/\partial S^{2})_{Q}$ as a function of the
coupling constant parameter $\alpha $ for different values of the size of
black brane $r_{+}$. In order to investigate the stability of black branes,
we plot both $(\partial ^{2}M/\partial S^{2})_{Q}$\ and $T$ \ in one single
figure for various values of $r_{+}$\ or $\alpha $. Of course, one should
note that in these figures, only the positive values of temperature
associated to the event horizon of the black branes, and negative values of
temperature belong to inner horizon which we are not interested in. We
discuss these figures for $k=0$ and $k=-1$, separately.
\textbf{$k=0$}:
As we discussed in \ the last section, small black brane with a radius
r_{+} $\ exist when $\alpha >\alpha _{\mathrm{ext}}$. Figures \ref{Fig3} and
\ref{Fig4} show that these small black branes are stable provided $\alpha
>\alpha _{\mathrm{crit}}$. On the other side, large black branes are stable
as one may see in Figs. \ref{Fig5} and \ref{Fig6}.
\textbf{$k=-1$}:
Figures \ref{Fig3} and \ref{Fig4} show that small black branes
exist only for medium values of $\alpha $\ ($\alpha
_{1\mathrm{ext}}<\alpha <\alpha _{2\mathrm{ext}}$), but they are
unstable. On the other side, large black branes are stable as one
may see in Figs. \ref{Fig5} and \ref{Fig6}.
\begin{figure}[tbp]
\epsfxsize=8cm \centerline{\epsffile{fig1F.eps}}
\caption{$F^{\mathrm{off}}$ versus $r_{+}$ for $b=0.2$, $n=4$,
$l=1$, $k=0$, $\protect\alpha =1.2$ and $T=0.42$.} \label{Fig1F}
\end{figure}
In order to confirm the stability analysis of the black branes,
one can also use the generalized free energy \cite{Myung}
\[
F^{\mathrm{off}}(b,\alpha ,r_{+},T)=M(b,\alpha ,r_{+})-TS_{+}(b,\alpha
,r_{+}),
\]
which applies to any value of $r_{+}$ with a fixed
temperature $T$. This off-shell free energy reduces to the on-shell free
energy at $T=T_+$:
\[
F=M-T_{+}S_{+},
\]
which is the Legendre transform of $M$ with respect to $S$. As an
example of using the stability analysis by off-shell free energy,
we plot $F^{\mathrm{off}}$ versus $r_+$ for $k=0$, $\alpha=1.2$
and $\alpha=4$. These are plotted in Figs. \ref{Fig1F} and \ref{Fig2F},
which show that the black brane is unstable for $\alpha=1.2$,
while it is stable for $\alpha=4$. These figures confirm the
stability analysis of Fig. \ref{Fig4}. One may also confirm other
stability analysis given in this section.
\begin{figure}[tbp]
\epsfxsize=8cm \centerline{\epsffile{fig2F.eps}}
\caption{$F^{\mathrm{off}}$ versus $r_{+}$ for $b=0.2$, $n=4$,
$l=1$, $k=0$, $\protect\alpha =4$ and $T=0.0001$.} \label{Fig2F}
\end{figure}
\section{Conclusions\label{Sum}}
In $(n+1)$-dimensions, when the $(n-1)$-sphere of black hole event
horizons is replaced by an $(n-1)$-dimensional hypersurface with
zero or negative constant curvature, the black hole is referred to
as a topological black brane. The construction and analysis of
these exotic black branes in AdS space is a subject of much recent
interest. This is primarily due to their relevance for the AdS/CFT
correspondence. In this paper, we further generalized these exotic
black brane solutions by including a dilaton and the
electrodynamic fields in the action. We obtained a new class of $(n+1)
-dimensional $(n\geq 3)$ topological black brane solutions in
Einstein-Maxwell-dilaton gravity in the background of AdS spaces. Indeed,
the dilaton potential plays a crucial role in the existence of these black
brane solutions, as the negative cosmological constant does in the
Einstein-Maxwell theory. In the absence of a dilaton field ($\alpha =\gamma
=0$), our solutions reduce to the $(n+1)$-dimensional topological black brane
solutions presented in \cite{Cai33}. We computed the entropy, temperature,
charge, mass, and electric potential of the topological dilaton black branes
and found that these quantities satisfy the first law of thermodynamics.
For the thermodynamical analysis of the black branes, we divide
the black branes into three classes. The black brane is said to be
\textquotedblright small\textquotedblright\ when the radius of
event horizon is much smaller than the parameter $b$ , it is
called \textquotedblright medium\textquotedblright\ if $r_{+}$\
and $b$\ have the same order and it is
called \textquotedblright large\textquotedblright\ for large values of
r_{+} $\ with respect to $b$. The radius of event horizon is always larger
or equal to $r_{\mathrm{ext}}$\ for the black branes with flat horizon,
while it is in the range $r_{_{+}}<r_{1\mathrm{ext}}$\ or $r_{+}>r_{2\mathrm
ext}} $\ for hyperbolic black branes. We found that the existence of $r_
\mathrm{ext}}$\ for $k=0$, and $r_{1\mathrm{ext}}$\ for $k=-1$\ depends on
the value of $\alpha $. For the case of $k=0$\ with large values of $\alpha
\ ($\alpha >\alpha _{\mathrm{ext}}$), one can have black brane with any
size. This feature shows that dilaton change the thermodynamics of black
branes drastically. In this case, the temperature goes to infinity as $r_{+}
\ goes to zero for $n>4$, while it approaches to a constant for $n=4$. For
the case of $k=-1$, one may have small black branes provided $\alpha _{
\mathrm{ext}}<\alpha <\alpha _{2\mathrm{ext}}$, while large black branes
exist for any value of $\alpha$. Again, this is a drastic change in the
properties of the solutions because of the dilaton field. We analyzed the
thermal stability of the solutions in the canonical ensemble by finding a
Smarr-type formula and considering $(\partial ^{2}M/\partial S^{2})_{Q}$\
for the charged topological dilaton black brane solutions in $(n+1)$\
dimensions. We showed that for the case of $k=0$, small black branes are
unstable for $\alpha _{\mathrm{ext}}<\alpha <\alpha _{\mathrm{crit}}$, while
they are stable for $\alpha >\alpha _{\mathrm{crit}}$. Also, in this case
large black branes are stable for arbitrary $\alpha$. For the case of $k=-1
, small black branes are unstable, while large ones are stable. That is
dilaton with the potential given in Eq. (\ref{V(Phi)}), changes the
stability of the black branes drastically.
\acknowledgments{We thank the referees for constructive comments.
This work has been supported by Research Institute for Astronomy
and Astrophysics of Maragha, Iran.}
|
\section*{Introduction}
Magnetostriction, the strain response of a material to an
external or internal magnetic field, is a thermodynamic tensor
quantity that is closely related to the magnetisation.\cite{doerr08}
Pulsed field facilities provide the highest available magnetic fields
for research purposes. Electrical and mechanical noise are
significantly higher than in steady fields, however, and the short
timescales mean that this noise cannot be effectively averaged. For a
typical pulse with a maximum field of $B=50$\,T occurring in $10$\,ms,
the field sweep rate can reach $\sim 5000$\,T/s. In this demanding
environment, fast optical techniques which are immune to
electromagnetic interference have significant advantages.
We first review the established methods of measuring magnetostriction in
pulsed fields, via capacitance dilatometry or resistive
foil strain gauges.
Capacitance dilatometry is the standard technique used for both
thermal expansion and magnetostriction measurements in steady
fields. The main advantages are that the sample is under only a weak
uniaxial stress (applied via springs to keep the capacitor plates in
contact with the sample), and that the sensitivity increases the
closer the capacitor plates can be approached, while kept in a
parallel configuration. Using this technique, the minimum relative
change in sample length, $L$, that can be resolved is $\Delta L/L \sim
10^{-5}$ at 10--20\,kHz in pulsed fields\cite{doerr08} (or $\sim
10^{-7}/\sqrt{\text{Hz}}$). Such dilatometers are sensitive to
vibration, since the plates are only coupled to each other through the
sample, and to electrical noise in the measurement circuit. The
limited space inside pulsed field magnets acts as a further constraint
on the sensitivity, as the capacitor plates cannot be above a certain size.
Resistive foil gauges bonded directly to the sample have the benefit
of relative immunity from vibrations, but suffer from strong
magnetoresistance which must be calibrated out or well matched. Their
sensitivity is also quite low. For a 1\,k$\Omega$ strain
gauge a resistance change of 2\,m$\Omega$ must be detected to
resolve strains of $10^{-6}$. The best resolution so far reported
with this technique is $5 \times 10^{-6}$ in pulsed
fields.\cite{zaragoza06} One further disadvantage is the need to
bond strain gauges to the sample surface. The resulting
inhomogeneous stress field can affect the response of softer
samples. On the other hand, this confers relative immunity from
vibration.
An electrically equivalent, but potentially more sensitive technique
is to use the the deflection of a piezoresistive cantilever to measure the
magnetostriction. This has been recently demonstrated in steady and
pulsed high fields.\cite{park09} Minimal stress is applied to the sample from contact
with the thin cantilever. The sensitivity is as yet unknown, however.
Common to all three techniques described above is the need for an
electrical bridge circuit to measure either resistance or
capacitance. The metallic components of the bridge circuit, such as
wiring and capacitor plates, are subject to self-heating and motion in
the rapidly varying magnetic field. In general, these cause both noise
and a background signal which can be difficult to subtract from the output.
In contrast, optical fibre strain gauges --- or Fibre Bragg Gratings
(FBGs) --- have the significant advantage of immunity to
electromagnetic interference. In addition they combine some of the key
advantages of the technologies described above. They can be bonded
directly to samples in the same way as resistive foil strain gauges,
or can be incorporated into a dilatometer design where they are used
both as spring and sensing element.
An FBG is based on a modulation of the refractive index in the core of
an optical fibre over some length. The period of this modulation
determines the wavelength of light reflected (the Bragg wavelength),
while the depth of modulation controls the reflectivity. Longer FBGs
have narrower linewidths.\cite{othonos97} FBGs have several
applications in the fields of telecommunications and sensors, and are
a quite mature (and hence cost effective and widely available)
technology.
In this article, we discuss the implementation of FBG strain gauge
dilatometry at cryogenic temperatures in pulsed high magnetic
fields and show that the resolution and interrogation rate are
limited only by mechanical noise and the measurement system
respectively. We discuss the ways in which strain can be transmitted
from the sample under investigation to the FBG. We describe a high
resolution, high speed measurement system capable of resolving changes
in strain of $\sim 10^{-9}/\sqrt{\text{Hz}}$ in pulsed fields.
\section*{Methods}
The Bragg wavelength, $\lambda_B$, of an FBG is related to the
refractive index of the optical fibre, $n$, and the pitch of the
grating, $\Lambda$:
\begin{equation}
\lambda_B = 2n\Lambda
\end{equation}
The strain and temperature dependence of $\lambda_B$ are given by:
\begin{equation}
\begin{split}
\frac{\Delta\lambda_B}{\lambda_B} &=
\biggl( 1 - \frac{n^2}{2}(P_{12} - \nu(P_{11}+P_{12}))\biggr)
\frac{\Delta L}{L} \\
&+ \biggl(\alpha(T) + \frac{1}{n}\frac{dn}{dT}\biggr)
\Delta T
\end{split}
\end{equation}
The strain response depends on the collection of terms
$1-\frac{n^2}{2}(P_{12} - \nu(P_{11}+P_{12})) \approx
0.76$,\cite{kersey97} which has minimal temperature dependence from
2--300\,K for FBGs in silica fibre.\cite{james02} $P_{11}$ and
$P_{12}$ are components of the strain-optic tensor and $\nu$ is the
Poisson ratio. The strain response of an FBG is thus linear with the
same constant of proportionality at all temperatures, and no
calibration is required.
The temperature dependence is controlled by the linear thermal
expansion of the optical fibre, $\alpha(T)$, and the change in
refractive index with temperature. Both of these become very small at
$^4$He temperatures in silica fibres.\cite{reid98}
The change in $\lambda_B$ in high magnetic fields has been calculated
for the case of polarised light,\cite{diego04} and would lead to a
shift of around 0.1\,nm for an FBG with $\lambda_B=$1550\,nm in silica
at 100\,T. For the case of unpolarised light, as is used in typical
interrogation systems, $\lambda_B$ is not expected to change with magnetic
field.
\begin{figure}
\includegraphics[width=0.47\textwidth]{daouFig1.eps}
\caption{{\em a)} Optical circuit used for
interrogation of the FBG. The boxes show schematic spectra of the
light at each point in the circuit. {\em b)} The reflection
spectrum for the FBG pictured above at 5.4\,K in
zero field (red squares). The r.m.s. noise in each pixel is shown multiplied by
100 (black line). {\em c)} Optical fibre containing 2\,mm FBG glued to a
single crystal of cubic GdSb with cyanoacrylate. The sample
dimensions are $3 \times 1.5 \times 1.5$\,mm$^3$. The sample is glued
to a flat surface perpendicular to the optical fibre and sample
$a$-axis using GE varnish. This arrangement is appropriate for
measuring $\Delta L || B || a$, where $B$ is the applied magnetic
field. The measurement probe on which the sample is mounted is
8\,mm in diameter, suitable for the smallest bore magnets. The
optical fibre runs to the top of the probe where it passes through
a vacuum feedthrough.}
\label{fig:glued}
\end{figure}
We have chosen to work with FBGs written into the core of single-mode
telecommunications optic fibre with $\lambda_B = 1550$\,nm
at room temperature. Since this is an important telecoms band,
interrogation equipment and custom FBGs are readily available. The
expected shift of $\lambda_B$ is 1.2\,pm for 10$^{-6}$ strain.
The fibre we use has a core diameter of 9\,$\mu$m and a cladding
diameter of 125\,$\mu$m. On top of this there is a protective plastic
coating. The choice of fibre coating is important for strain-related
applications. The mechanical contact between sensor and sample must
accurately transmit the strain. Polyimide is a good choice as it
sticks to the fibre strongly and there are many epoxies suitable for
fixing the FBGs to the test material. Including coating, the FBGs that
we use are $\sim$0.16\,mm in diameter and either 1 or 2\,mm long. This
is smaller than the majority of commercially available resistive foil
strain gauges. A maximum strain of around $10^{-2}$ can be applied to
an FBG before the response becomes non-linear or the fibre
breaks.\cite{smartfibres} This is easily large enough for most strain
measurements.
There are then many choices of interrogation systems to use,
operating on diverse principles. For a good review, see
Ref.~\onlinecite{kersey97}. For measurements in pulsed fields a fast
technique ($>$\,10kHz) is necessary, whereas absolute accuracy and
long-term stability are less important.
We focus our attention on multi-channel spectrometry, which allows us
to detect the complete lineshape of the FBG reflection. If the
lineshape is distorted when we affix the FBG to the sample, we can
tell if the FBG is under homogeneous stress, {\em i.e.} if it is
uniformly coupled to the sample along its length.
Our high-resolution FBG interrogation system uses the fastest
currently available 1024-pixel InGaAs line array camera sensitive in
the wavelength range 900 to 1700\,nm, which can read full spectra at
47\,kHz.\cite{sensorsunlimited} The signal-to-noise ratio (SNR) is
specified at 5300:1 at full scale where readout and shot noise
dominate. Mounted on a fibre-coupled research spectrometer of focal
length 550\,mm with a 950\,l/mm diffraction grating, the 25\,$\mu$m
pixel pitch results in a dispersion, $\Delta\lambda$, of
0.027\,nm/pixel at 1550\,nm.
To illuminate the FBG we use a superluminescent-diode-based
broadband source with an output power of 60\,mW in the wavelength
range 1525-1565\,nm.\cite{denselight} The optical circuit is illustrated in
Figure~\ref{fig:glued}a. When the source characteristic is measured
directly using the spectrometer and camera described above, the SNR
reaches 800:1 for full-scale illumination at the shortest integration
time of 20.4\,$\mu$s. This is the dominant source of noise in the measurement.
A typical reflection spectrum for a 2\,mm FBG is shown in
Figure~\ref{fig:glued}b. The uncertainty in the measurement of
$\lambda_B$ depends on the way it is calculated from the returned
spectra. The standard technique is to calculate the centre of mass of
the reflected peak using:
\begin{equation}
\lambda_B =\frac{\Sigma_i \lambda_i S_i}{\Sigma_i S_i}
\label{eqn:lambdab}
\end{equation}
where $\lambda_i$ is the wavelength at the centre of the $i$-th pixel
and $S_i$ is the corresponding signal amplitude. For a peak whose full
width covers $m$ pixels, the r.m.s. error in $\lambda_B$ is
approximately:
\begin{equation}
\sigma_{\lambda_B} \approx \frac{m^{1/2}\Delta\lambda}{6\mathrm{ SNR}} \qquad (m \gg 1)
\label{eqn:error}
\end{equation}
which has been obtained by a simulation excluding the effects of
camera readout noise. Our 2\,mm FBGs have a full width at half-maximum
(FWHM) of $1.0 \pm 0.1$\,nm and $m\sim 80$. For the parameters quoted
above, this should result in a strain resolution of $\Delta L/L \sim 5
\times 10^{-8}$ per reading, or $\sim 2\times
10^{-10}/\sqrt{\text{Hz}}$.
One pitfall of this approach is that the incident light power must be
increased as the rate of interrogation increases, since the detector
must be saturated in a shorter time to achieve the optimal
SNR. Fortunately, this power is not dissipated at the sample, rather
it is lost at the end of the fibre. If the cryostat is wider than the
minimum bend diameter of the fibre (10\,mm), its end can be positioned
back at the warm end of the cryostat where extra heat input is
irrelevant. This means that FBGs can be interrogated at very low
temperatures without self-heating, making them suitable for operation
in $^3$He or dilution refrigerators. This is an improvement on
electrical measurements, where the excitation must be reduced at low
temperatures to avoid self heating, which necessarily reduces the SNR.
Coupling the FBG to the sample can be done in two ways. Our primary
consideration is that the experiment should fit inside the various
cryostats used for pulsed field measurements, the smallest of which
has an internal diameter of 8\,mm.
In the simplest case, for a regular sample with a flat surface parallel to
the direction of interest, the FBG can be bonded directly to the
sample using an appropriate epoxy.\cite{hbmepoxy} The considerations
are the same as for resisitive foil strain gauges: what curing
conditions (temperature and pressure) can be tolerated by the sample and FBG,
how good is the bond, how accurately can the FBG be aligned along a
crystalline axis? Of particular concern is that the sample can
experience an inhomogeneous stress due to the epoxy, which therefore
distorts the results.
To estimate if this is significant, consider that the Young's modulus
of silica is around 70\,GPa. This is similar to many metals. As long
as the sample has a much larger cross-sectional area than the optical
fibre, the presence of the fibre should not have a significant effect.
The advantages of direct bonding of the FBG to the sample surface are
compactness and less sensitivity to mechanical vibrations, which are
both concerns when working in small bore pulsed magnets. We have also
substituted cyanoacrylate (superglue) in place of a specialised epoxy,
with some success. This cures quickly at room temperature and is
soluble in acetone, allowing both sample and FBG to be
recovered. Figure~\ref{fig:glued}c shows a sample with an FBG glued
along its length mounted on a measurement probe.
The alternative is to couple the FBG and sample indirectly, through a
dilatometer. In such a dilatometer the sample is clamped into a
holder which is mechanically coupled to a free standing FBG sensor. In
this way the strain is directly transmitted to the sensor. This is
preferable for samples that are intolerant of applied stress, for
irregularly shaped samples, or for those which present a different
behaviour on the surface than in the bulk.
\section*{Results}
We have tested the FBG interrogation system described above in
pulsed magnetic fields at the Hochfeld-Magnetlabor Dresden.
We present results on single-crystal GdSb and GdSi. Owing to the lack of
crystal field effects ($L$~=~0 for Gd$^{3+}$), these two compounds are model
systems in which to study complex magnetic two-ion interactions. Both
reveal large field-induced magnetostriction.
Figure~\ref{fig:gdsb} shows the response of the FBG on the GdSb single
crystal pictured in Fig.~\ref{fig:glued}c when exposed to the pulsed
magnetic field. GdSb is a member of the Gd monopnictide series Gd$X$
($X$=N, P, As, Sb and Bi) with a cubic NaCl-type crystal structure and
is characterized by a low carrier density. It orders
antiferromagnetically at 23\,K and has nearly no magnetic
anisotropy. The transition into the field-induced ferromagnetic state
proceeds over a wide field range up to 32.6\,T followed by a sharp kink
when saturation is reached. The magnetostriction measurement shows
this transition clearly. Despite the strong spin moment of
Gd$^{3+}$, the change in the magnetic moment is not accompanied by a
large lattice effect. The magnetostriction is in the order of
$10^{-4}$ only.
The r.m.s. variation of $\lambda_B$ over 15\,ms in zero field is shown
in Figure~\ref{fig:gdsb}a
to be $\sigma_{\lambda_B} = 0.18$\,pm, which corresponds to a strain
resolution of $1.5 \times 10^{-7}$. This is three times less than
the (optimistic) prediction of Equation~\ref{eqn:error}, and we
attribute the difference to a combination of camera readout noise and
mechanical vibration.
The noise at high fields is similar to that at zero field. There is
some evidence that the sample is heated by the rapidly changing
magnetic field, as the value of $\lambda_B$ immediately after the
pulse is slightly different to that before. This can be seen in
Figure~\ref{fig:gdsb}b. Since the thermal expansion intrinsic
to the optical fibre is negligble below 40\,K, it does not contribute
significantly to this shift.
The up- and downsweep data match extremely well above 20\,T when
plotted as a function of magnetic field Figure~\ref{fig:gdsb}c. Hence
any self-heating effect must be limited to long times or low fields.
More noise can be seen towards the end of the pulse (see also the
inset to Figure~\ref{fig:gdsi}). This may be due to mechanical
vibrations caused by the rapid boiloff of the liquid cryogens in the
system during the pulse.
\begin{figure}
\includegraphics[width=0.47\textwidth,clip]{daouFig2.eps}
\caption{{\em a)} $\Delta\lambda_B$ over time in zero field as
determined by Equation~\ref{eqn:lambdab}, for the FBG glued to the
single crystal of GdSb at 5.4\,K, as shown in
Figure~\ref{fig:glued}c (red line). The two horizontal lines are at
$\pm \sigma_{\lambda_B} = 0.18$\,pm. {\em b)} When exposed
to the magnetic field (blue line), $\Delta\lambda_B$ recovers to
nearly the same value after the field returns to zero. {\em
c)} The relative length change $\Delta L/L$ as a function
of magnetic field determined from the data above. The field sweep
direction is indicated by arrows. The increase below 2\,T is a
spin-reorientation out of the normal antiferromagnetic state,
while the sharp kink at $32.62 \pm 0.03$\,T corresponds to the
point where the spins are fully ferromagnetically aligned. The up
and down sweeps match extremely well at high fields. The signal
is noisier as the field returns to zero where mechanical vibrations
become stronger. The inset zooms in on the data above 30\,T, where
small variations in $\Delta L/L$ at the 10$^{-6}$ scale are very
well reproduced as field goes up and down, suggesting they are an
intrinsic response of the sample. The random noise at high fields
is around the same as in zero field.}
\label{fig:gdsb}
\end{figure}
In order to compare the performance FBG-based magnetostriction
measurements directly with established techniques, we have measured the same
single crystal of GdSi as was used in previous studies.
Figure~\ref{fig:gdsi} shows the magnetostriction of a single crystal
of GdSi measured by bonding a 1\,mm FBG to the sample with
cyanoacrylate. GdSi crystallizes in the orthorhombic FeB
structure. The magnetisation at low field shows anisotropic behaviour
due to the interplay between the localised $4f$-spins and the
conduction electrons.\cite{tung05} The magnetisation along the
$c$-axis merges with that measured along $b$ at about 3\,T. Note, that
this small anomaly is clearly visible in the magnetostriction curve
measured by FBG. At higher field the magnetisation becomes isotropic
and, after a spin-flop process, reaches saturation through a
first-order-like transition at 20\,T. Both processes are clearly
visible in the striction curve. The r.m.s. noise level in the GdSi
experiment is increased to about $\sigma_{\lambda_B} = 0.3$\,pm as the
shorter FBG has a wider reflection peak ($m\sim 160$). The shorter FBG
was required as the sample was less than 2\,mm long along the
$c$-axis.
The comparison of the FBG technique with high steady and pulsed field
capacitance dilatometry shown in Figure~\ref{fig:gdsi} is very
favourable. The high bandwidth allows sharp features to be easily
resolved, even on the fast upsweep, while the noise level approaches
the steady field results. The FBG signal at 4.8\,K is about 5\% smaller than the steady
field measurements at 4\,K. Assuming that the steady field measurements are authoritative, this
suggests that the transmission of strain from the sample to the FBG is
at worst 95\%, which is extremely good. The other possibility is that
the slight difference in sample temperature leads to this small
quantitative difference in the magnetostriction.
\begin{figure}
\includegraphics[width=0.47\textwidth,clip]{daouFig3.eps}
\caption{Relative length change $\Delta L/L$ of GdSi at 4.8\,K
measured by 1\,mm FBG glued to the sample surface with
cyanoacrylate (red solid line) as a function of magnetic field
for $\Delta L || B || c$. The acquisition rate was 30\,kHz. The
data is nearly identical to that taken in a 33\,T resisitive
magnet on the same sample at 4\,K (black dotted line), which has been
scaled by +5\% and offset to match the downsweep of the FBG
data. The inset shows that the small transition of amplitude $\sim
10^{-5}$ at 2.5\,T is clear in both measurements, indicating the
quality of the pulsed field data. For comparison, data taken by
capacitance dilatometry at 5\,K with an excitation frequencz of 10\,kHz (blue dashed
line)\cite{doerr08} on the same sample but with $\Delta L ||
B || a$, is noisier and has a large drift at high field where
no further length change is expected as the magnetisation is saturated.}
\label{fig:gdsi}
\end{figure}
\section*{Conclusions}
We have demonstrated that FBGs are a very suitable technology for
high-accuracy magnetoelastic investigations in small samples
under the many simultaneous strong constraints imposed by low
temperatures and the highest avalable magnetic fields.
We have developed an FBG interrogation system capable of determining
$\lambda_B$ with an r.m.s. error $\sigma_{\lambda_B}=0.18$\,pm and a
bandwidth of 47\,kHz. We have used this to measure the
magnetostriction of small samples glued straightforwardly to
FBGs, at cryogenic temperatures in the hostile environment generated
by pulsed magnetic fields. We are capable of resolving relative length
changes, $\Delta L/L$, as small as $3\times 10^{-7}$ (twice the
r.m.s. error in interrogation) on a timescale of 20\,$\mu$s, or
$1.4\times 10^{-9}/\sqrt{\text{Hz}}$. This is nearly two orders of magnitude
better than was previously possible by other techniques in
non-destructive high magnetic fields. This increase in resolution
greatly broadens the range of materials whose magnetostriction
can be investigated in pulsed magnetic fields.
The FBG based technique has the significant advantage of immunity from
electrical noise, and is furthermore quite simple to implement. Since
there are no metallic parts, there are no self-heating effects in the
magnetic field. No calibration is required. We have shown that the strain is well
transmitted from the sample to the FBG.
It should therefore be easy to extend the technique to measure strain
transverse to the applied field as well. This can be done
simultaneously with a longitudinal measurement using a single fibre
containing two FBGs with different $\lambda_B$ glued to the sample in
different orientations. No further equipment is required. Similarly,
multiple samples can be measured simultaneously using a single fibre,
the only limitations being space in the cryostat at the centre of the
field and the spectral range of the spectrometer.
The technique is also compatible with very low temperatures
($<$1\,K), as in principle no heat is dissipated at the cold end
of the cryostat. Inserting the optical fibre into a high-pressure
piston-cylinder type clamp cell to measure magnetostriction under
pressure in pulsed fields is also a promising possibility. Much
potential exists to extend the utility of FBGs further.
\begin{acknowledgments}
This work was supported by the MPG Research Initiative ``Materials
Science and Condensed Matter Research at the Hochfeld-Magnetlabor
Dresden''. We would like to thank M.~Bartkowiak and S.W.~James for useful discussions and
F.~Tr\"{u}ller for technical support.
\end{acknowledgments}
|
\section{Introduction \label{sec:Introduction}}
One of the main features of higher spin theories is that apparently the only way of building a nonabelian interacting theory is to consider an infinite set of fields with unbounded
value of the spin. Such an interacting theory is by now well-known: Vasiliev's equations \cite{Vasiliev:1990en,Vasiliev:1992av,Vasiliev:2003ev}.
These equations admit (anti) de Sitter spacetime $(A)dS$ as exact solution,
but not the flat limit case with vanishing cosmological constant.
The theory \cite{Vasiliev:1990en,Vasiliev:1992av} can be given a Lagrangian formulation \cite{BS1}, albeit of a non-standard type.
It has been proved recently \cite{Boulanger:2008tg} that an interacting nonabelian theory built as a perturbative deformation of the free Fronsdal theory in $(A)dS$ spacetime \cite{Fronsdal:1979vb} and containing
an infinite tower of totally symmetric tensor gauge fields with unbounded
spin, does not admit any consistent flat limit. There are thus doubts about the mere existence of a nonabelian theory around Minkoswki spacetime.
To emphasize this, we show in this paper that a standard requirement about
higher spin nonabelian interactions in Minkowski spacetime cannot, at least
in dimension four, solve the usual flat-space interaction problems.
Some years ago, Berends, Burgers and van Dam (BBvD) exhibited a pure spin-3 nonabelian cubic vertex in flat spacetime \cite{Berends:1984wp}, and then found that this vertex cannot be further extended to higher orders in deformations
if only spin-3 fields are considered in the spectrum of fields \cite{Berends:1984rq}. A similar result had also been obtained in \cite{Bengtsson:1983bp}.
In more technical terms,
an obstruction to the existence of pure spin-3 quartic deformations appeared.
In the light of this negative result, they brought the idea that the obstruction, as well as others of the same kind, could probably be cured by introducing fields of spin higher than 3. Recursively, this would suggest
that every value of the spin is needed in order to build a nonabelian higher spin theory in flat spacetime.
In this paper, we actually prove that the BBvD vertex is in fact strongly obstructed in dimension strictly higher than three, in the following sense:
Even upon introducing higher (and/or lower) spin gauge fields, it is not possible to cure the obstruction brought in by the spin-3 BBvD vertex found
in \cite{Berends:1984wp}.
The very reason for this strong obstruction in flat background is that, as opposed to what happens in $(A)dS$ background, the number of derivatives involved in an expression constitutes a well-defined grading that increases with the value of the spin.
In $(A)dS$ instead, the non-zero commutators of covariant derivatives introduce expansions in powers of the cosmological constant involving different numbers of derivatives. Furthermore, these expansions in powers of the cosmological constant are precisely what prevents one from considering a consistent nonabelian flat limit of the $(A)dS$ theories, as was mentioned
already in \cite{Fradkin:1987ks,Fradkin:1987qy}.
The paper is organized as follows. In Section \ref{sec:af} we review Fronsdal's theory in the antifield formulation, as well as the cohomological reformulation of the consistent deformation problem.
In Section \ref{sec:genr} we present our theorem on the lowest-order deformations of the gauge algebra arising from local cubic interactions between totally symmetric higher spin gauge fields in flat spacetime. This is used in Section \ref{sec:obstr} in order to address the specific case of the BBvD vertex. In Section \ref{concl}, these results are briefly summarized as a conclusion.
\section{Antifield formulation and consistent deformations \label{sec:af}}
We use the antifield formalism \cite{Henneaux:1992ig,Barnich:1993vg,Henneaux:1998i,Barnich:1995db} to derive our results about consistent deformations of Fronsdal Lagrangians \cite{Fronsdal:1978rb}. The BBvD vertex only involves the spin-3 gauge fields, but since we allow them to mix with other gauge fields, we recall the Fronsdal Lagrangian for an arbitrary spin gauge field in flat spacetime, as well as the corresponding antifield formulation.
\subsection{The Fronsdal Lagrangian}
The field denoted by $\phi^a_{\mu_1...\mu_{s_a}}$ is a totally symmetric field of spin $s_a$ and double-traceless : $\eta^{\mu_1\mu_2}\eta^{\mu_3\mu_4}\phi^a_{\mu_1...\mu_{s_a}}\equiv 0\,$.
The index $a$ labels a given set of fields with various values of the spin.
The Fronsdal tensor reads:
\begin{eqnarray}
F^a_{\mu_1...\mu_{s_a}} := \Box \phi^a_{\mu_1...\mu_{s_a}}-s_a\,\partial^{\rho}\partial^{}_{(\mu_1}\phi^a_{\mu_2...\mu_{s_a})\rho}+\frac{s_a(s_a-1)}{2}\;\partial^{}_{(\mu_1}\partial^{}_{\mu_2}\phi'{}^a_{\mu_3...\mu_{s_a})}\quad,
\end{eqnarray}
where $\phi'{}^a$ stands for the trace of $\phi^a\,$.
The Fronsdal tensor is invariant under the gauge transformations:
\begin{eqnarray}
\delta_\xi\phi^a_{\mu_1...\mu_{s_a}}=s_a\,\partial^{}_{(\mu_1}\xi^a_{\mu_2...\mu_{s_a})}\quad,\label{gts}
\end{eqnarray}
where the gauge parameter $\xi^a$ is traceless.
The generalized Einstein tensor is defined as:
\begin{eqnarray}
G^a_{\mu_1...\mu_{s_a}} := F^a_{\mu_1...\mu_{s_a}} - \frac{s_a(s_a-1)}{4}\;\eta^{}_{(\mu_1\mu_2}F'{}^a_{\mu_3...\mu_{s_a})}
\quad.
\end{eqnarray}
Finally, the Lagrangian can be written: \begin{eqnarray}{\cal{L}}_F=\frac{1}{2}\sum_a \phi^{\mu_1...\mu_{s_a}}_a G^a_{\mu_1...\mu_{s_a}}\quad.\end{eqnarray}
\subsection{Antifield formulation}
A set of fermionic ghosts $C_{\mu_1...\mu_{s_a-1}}^a$ is introduced, with the same tensorial
structure as the associated gauge parameter. In particular, the ghosts are traceless.
They carry a pure ghost number 1 (also denoted $pgh\ C^a=1$). Then, two families of antifields are associated with the fields and the ghosts : the fermionic antifields $\phi^{*\mu_1...\mu_{s_a}}_a$ and the bosonic antifields $C^{*\mu_1...\mu_{s_a-1}}_a$. The antifield number (denoted $antigh$) counts the number of antifields with the following weight: $antigh\ \phi^*_a=1$ and $antigh\ C^*_a=2$.
The longitudinal derivative $\gamma$ has a vanishing action on every field except $\phi^a$, for which:
\begin{eqnarray}
\gamma \phi^a_{\mu_1...\mu_{s_a}} = s_a\, \partial^{}_{(\mu_1}C^a_{\mu_2...\mu_{s_a})}\quad.
\end{eqnarray}
On the other hand the Kozsul--Tate differential has a non-vanishing action only on the antifields: \begin{eqnarray}
\delta\phi^{*\mu_1...\mu_{s_a}}_a &:=& \frac{\delta {\cal{L}}_F}{\delta \phi^a_{\mu_1...\mu_{s_a}}}
~~\textrm{ and }
\nonumber \\
\delta C^{*\mu_1...\mu_{s_a-1}}_a &:=&
-s_a\,\partial^{}_\rho\left[\phi^{*\mu_1...\mu_{s_a-1}\rho}_a-\frac{(s-1)(s-2)}{n+2s-6}\;\eta^{(\mu_1\mu_2}
\phi_a^*{}'{}^{\mu_3...\mu_{s_a-1})\rho}\right]\quad ,
\end{eqnarray}
where $n$ denotes the dimension of the flat spactime.
\noindent The generator $\stackrel{(0)}{W}$, also called ``solution of the master equation'', is then introduced:
\begin{eqnarray}
\stackrel{(0)}{W}\;\,=\int\left[{\cal{L}}_F + s_a\,\sum_a\phi^{*\mu_1...\mu_{s_a}}_a\partial_{\mu_1}C^a_{\mu_2...\mu_{s_a}}\right]d^n x\quad.
\end{eqnarray}
Let us define the antibracket (we denote collectively the fields and ghosts as $\Phi^i$ and the antifields as $\Phi^*_i$):
\begin{eqnarray}
(A,B)=\frac{\delta^L A}{\delta\Phi^i}\frac{\delta^R B}{\delta\Phi^*_i} -
\frac{\delta^L A}{\delta\Phi^*_i}\frac{\delta^R B}{\delta\Phi^i}\quad.
\end{eqnarray}
The generator satisfies: $\Big(\stackrel{(0)}{W},A\Big) = {\textbf{s}} A$ where ${\textbf{s}}=\delta+\gamma$ is the BRST differential of the theory.
\subsection{Cohomology of $\gamma$}
In this section, we introduce our notation for the cohomology of $\gamma$ (whose elements are called \textit{invariants}). It has been showed \cite{Bekaert:2005ka} that the local functions of $H^*(\gamma)$ for a spin-$s_a$ Fronsdal theory in flat spacetime only depend on the antifields, the Fronsdal tensor $F^a_{\mu_1...\mu_{s_a}}$, the curvature tensor $K^a_{\mu_1\nu_1|...|\mu_{s_a}\nu_{s_a}}$ (which consists of $s_a$
curls of the field) and their derivatives, as well as some non $\gamma$-exact ghost tensors, denoted $U^{(i)a}_{\mu_1\nu_1|...|\mu_i\nu_i|\nu_{i+1}...\nu_{s_a-1}}$ ($i<s_a$), that are the traceless part of the $i$ times antisymmetrised $i$th derivatives of the ghosts. For example ($i=1$): \begin{eqnarray}U^{(1)a}_{\mu_1\nu_1|\nu_2...\nu_{s_a-1}}=\partial_{[\mu_1}C^a_{\nu_1]\nu_2...\nu_{s_a-1}}-\frac{(s-2)}{n+2s-4}\eta^{}_{[\mu_1|(\nu_2}\partial^\rho C^a_{\nu_3...\nu_{s_a-1})|\nu_1]\rho}\quad.\end{eqnarray}
Of course, the zeroth tensor ($i=0$) is the undifferentiated ghost itself.
More generally, the ghost tensors $U^{(i)a}_{\mu_1\nu_1|...|\mu_i\nu_i|\nu_{i+1}...\nu_{s_a-1}}$ give irreducible representations of the Lorentz algebra $\mathfrak{o}(n-1,1)$ labeled by Young diagrams made of two rows of respective lengths $s_a-1$ and $i<s_a$. This property will be extremely useful in order to classify the nonabelian cubic deformations.
In the case of a sum of Fronsdal theories, the cohomology of $\gamma$ is simply the direct product of the cohomologies of the separate theories. The cohomology is thus the following set of functions: \begin{eqnarray}H^*(\gamma)=\left\{f([\Phi^*_a],[F^a],[K^a],C^a,U^{(i)a})\right\}\quad,\end{eqnarray} where the square brackets around a field here denotes the corresponding field and all its derivatives. The functions $f$ are polynomials in the local case (and the number of derivatives is bounded from above).
\subsection{Consistent deformations}
The problem of consistently deforming a free theory, like Fronsdal's theory, into a full, interacting theory, can be reformulated within the antifield formalism \cite{Barnich:1993vg}: if one considers an expansion of the generator $W$ in terms of a parameter $g$:
$W = \int w \,= \;\stackrel{(0)}{W}+\,g\stackrel{(1)}{W}+\,g^2\stackrel{(2)}{W}+\,...\,$, then the deformation is consistent if the full generator satisfies the {\textit{master equation}} $(W,W)=0$ to all orders in $g$. Since Fronsdal's theory is consistent, the initial generator satisfies $(\stackrel{(0)}{W},\stackrel{(0)}{W})=0$, which implies that ${\textbf{s}}$ is a differential (${\textbf{s}}^2=0$).
The first order equation is $(\stackrel{(0)}{W},\stackrel{(1)}{W})={\textbf{s}}\stackrel{(1)}{W}\,=0\,$. In the case of a local deformation, $\stackrel{(1)}{W}$ must be the integral of a local $n$-form $a:=\,\stackrel{(1)}{w}$, and the equation
${\textbf{s}}\stackrel{(1)}{W}\,=0$ becomes a ${\textbf{s}}$-cocycle relation modulo ${\rm d}$:
\begin{eqnarray}{\textbf{s}} \,a\,+\,{\rm d} \,b=0\quad,
\end{eqnarray}
where the operator ${\rm d}$ denotes the total exterior differential.
Since ${\textbf{s}}$-exact and ${\rm d}$-exact terms in the cocycle $\stackrel{(1)}{w}$ correspond to trivial deformations, the first order inequivalent deformations are described by the cohomology class $H^{0,n}({\textbf{s}}|{\rm d})\,$,
see \cite{Barnich:1993vg,Henneaux:1998i}.
The obstructions under consideration in this paper arise when checking whether the first order vertices satisfy the master equation at second order, the local form of which reads:
\begin{eqnarray}(\stackrel{(1)}{w},\stackrel{(1)}{w})\,d^n x=-\frac{1}{2}\,{\textbf{s}}\stackrel{(2)}{w}+\,{\rm d}\,e
\label{soe}\quad.
\end{eqnarray}
\subsection{Cubic vertices}
It has been showed, for values of the spin up to 4, that the only first order nonabelian solutions of the local master equation in flat spacetime are cubic in the fields, ghosts or antifields \cite{Barnich:1994,Boulanger:2000rq,Bekaert:2006jf,Boulanger:2008tg}. For the spin-2 fields, this allows one to show \cite{Boulanger:2000rq} that Einstein's gravity is the only nonabelian consistent deformation with at most two derivatives of
the free Pauli-Fierz theory in dimension $n>3\,$. For the spin-3 case, we
know \cite{Bekaert:2006jf} that the BBvD vertex is one of the only two possible nonabelian first order deformations of the spin-3 Fronsdal theory.
For values of the spin strictly greater than 4, it is still not proved
within the antifield formalism whether other kinds of deformations are possible, for example starting with a quartic first order vertex. However, the problem consisting in computing cubic first order deformations can be addressed. As will be shown in the next subsection, the classification of the candidates is severe enough to put strong constraints on the allowed number of derivatives and on the gauge structure of the deformation, depending on the spins involved. Let us also mention the very powerful light-cone gauge method
used in~\cite{Metsaev:2005ar,Metsaev:2007rn}.
The components of the generator $w$ carry a ghost number $0$, which means that their antifield number is equal to their pure ghost number.
Hence, for cubic deformations, only expressions with at most antifield number 2 will appear, since the pure ghost number of any individual field is at most 1
in Fronsdal's theory.
Therefore the cubic first order deformation can be
expanded in the antifield number $a=a_0+a_1+a_2$ and the master equation decomposes into the
following system of equations:
\begin{eqnarray}
&\gamma \,a_2=0\label{eqagh2}& \\
&\delta \,a_2+\gamma \,a_1+{\rm d }\, b_1=0\label{eqagh1}&\\
&\delta \,a_1+\gamma \,a_0+{\rm d }\, b_0=0\label{eqagh0}&\quad.
\end{eqnarray}
The component $a_2$ contains the information about the first order deformation of the
({\textit{a priori}} on-shell) gauge algebra.
The nonabelian deformations are thus characterized by a non-vanishing $a_2$ component.
A cubic $a_2$ is linear in the antifield number $2$ antifields and quadratic in the ghosts, it does not depend on the fields.
Consequently, the gauge algebra closes off-shell at first-order in the deformation, for cubic vertices. Moreover, the top form $a_2$
is $\gamma$-closed by Eq.(\ref{eqagh2}) and any $\gamma$-exact term is trivial in the sense that it is the antifield number $2$ part of an ${\textbf{s}}$-exact term in $a$. Thus $a_2$ can efficiently be written as a representative of $H^2(\gamma)$. Finally, $a_2$ is defined modulo ${\rm d }\,$, which allows one to only consider undifferentiated antifields. Therefore, without loss of generality, the general structure of $a_2$ reads schematically:
\begin{eqnarray}
a_2=f^a_{bc|(i)(j)}\,C^{*}_a\, U^{(i)b}\, U^{(j)c}\,d^n x\quad,
\end{eqnarray}
where $f^a_{bc|(i)(j)}$ are internal coefficients. This expression of $a_2$ encodes the structure constants of the gauge algebra, at first order in the deformation. A Poincar\'e invariant $a_2$ is Lorentz-invariant (the spacetime indices must all be contracted) and does not explicitly depend on the spacetime coordinates, therefore the coefficients $f^a_{bc|(i)(j)}$ are constants.
Finally, the component of maximal antifield number of the second order equation (\ref{soe}),
which is the test that we use to exhibit the obstructions, reads:
\begin{eqnarray}(a_2,a_2)\,d^n x = \gamma \,c_2+{\rm d }\, e_2\quad.
\label{secondorder}
\end{eqnarray}
This equation is the translation, within the antifield formalism, of the
lowest-order component of the Jacobi identity for the gauge algebra.
\section{General results on the gauge algebra deformations \label{sec:genr}}
In this section, we provide general arguments that simplify the classification
of the cubic nonabelian deformations for an arbitrary spin configuration. For
any cubic configuration of the type $s-s'-s''$ (\textit{i.e.} including fields
of respective spins $s$, $s'$ and $s''$), with $s\leqslant s'\leqslant s''$,
there is a small number of possibilities of building consistent $a_2$
expressions, and only some of them are related to a consistent $a_1\,$.
As we said previously, a cubic $a_2$ can always be written in the form
$$a_2=C^* U^{(i)}\, U^{(j)}d^n x+\gamma(...)\quad,$$
where the $U^{(i)}$ are non-$\gamma$-exact ghost tensors.
A strong constraint on such candidates $a_2$ is that the product
$U^{(i)} U^{(j)}$ of ghost tensors (in general, with implicit contractions of
indices) must be contracted with the antighost $C^*$ which is itself a
symmetric Lorentz tensor. The Littlewood-Richardson rules will be used
throughout this section in order to analyze all possible contractions of the
indices from the two ghost tensors and the antifield.\footnote{More precisely,
the specific rules for product and division of Young diagrams are applied here
(see \textit{e.g.} the appendix A of \cite{Bekaert:2005ka} for a self-contained
review of these Littlewood-Richardson rules).} With the help of these rules, we
will show that the previous constraint implies several strong conditions on the
allowed values of the numbers of derivatives $i$, $j$ and of spins $s$, $s'$
and $s''\,$.
\subsection{Product of ghost tensors}\label{product}
Let us consider a product of two ghost tensors $U^{(i)}$ and $U^{(j)}$, corresponding respectively to spin $s_1$ and $s_2$ (with $s_1\leqslant s_2$).
\begin{enumerate}
\item[\textbf{A.}] Firstly, we may study the minimal number of free indices in that product (in other words, the maximal number of contracted indices).
\begin{itemize}
\item[\textbf{A.1.}] If $i\leqslant j$, all of the indices of
$U^{(i)}$ can be contracted with $s_1+i-1$ indices of $U^{(j)}$. Let us
visualize in terms of Young diagrams the symmetry properties of the tensors
resulting from the maximal contraction of indices:
\begin{eqnarray}U^{(i)}: \begin{picture}(60,0)(0,0)\multiframe(0,0)(10,0){1}(50,10){$s_1-1$}\multiframe(0,-10.5)(10,0){1}(20,10){$i$}\end{picture},\ U^{(j)}:\begin{picture}(70,0)(0,0)\multiframe(0,0)(10,0){1}(60,10){$s_2-1$}\multiframe(0,-10.5)(10,0){1}(40,10){$j$}\end{picture}\nonumber\end{eqnarray}\begin{eqnarray} &\Rightarrow {\textrm{Maximal contraction}}:&\bullet\ \begin{picture}(36,0)(0,0)\multiframe(0,-2)(10,0){1}(35,10){$s_2-s_1$}\end{picture}\otimes\begin{picture}(30,0)(0,0)\multiframe(0,-2)(10,0){1}(30,10){$j-i$}\end{picture}\textrm{ if $j<s_1$}\nonumber\\&&\bullet\ \bigoplus_a\begin{picture}(105,0)(0,0)\multiframe(0,1)(10,0){1}(100,10){$s_2-s_1+j-i-a$}\multiframe(0,-9.5)(10,0){1}(20,10){$a$}\end{picture}\textrm{ if $j-i\geqslant a\geqslant j-s_1+1>0$ }\nonumber\quad.\end{eqnarray}
Since $U^{(j)}$ bears $s_2+j-1$ indices, the minimal number $N_{min}$ of
free indices is $s_2-s_1+j-i$.
Furthermore, these free indices can be symmetrized if $j<s_1$ since there
is a component \begin{picture}(85,0)(0,0)\multiframe(0,-2)(10,0){1}(80,10){$s_2-s_1+j-i$}\end{picture}
in the tensor product. If $j\geqslant s_1$, no contraction of the two
tensors $U$ can be symmetrized and thus no Lorentz invariant $a_2$ can be
built. Consequently,
$$\mbox{max}\{i,j\}<s_1$$ in order to have symmetrizable free indices, as
can be seen for the other case as well.
\item[\textbf{A.2.}]
If $j<i<s_1\leqslant s_2$, let us visualize the ghost tensors:
\begin{eqnarray}U^{(i)}: \begin{picture}(80,0)(0,0)\multiframe(0,0)(10,0){1}(75,10){$s_1-1$}\multiframe(0,-10.5)(10,0){1}(20,10){$j$}\multiframe(20.5,-10.5)(10,0){1}(40,10){$i-j$}\end{picture},\ U^{(j)}:\begin{picture}(120,0)(0,0)\multiframe(0,0)(10,0){1}(75,10){$s_1-1$}\multiframe(75.5,0)(10,0){1}(40,10){$s_2-s_1$}\multiframe(0,-10.5)(10,0){1}(20,10){$j$}\end{picture}\nonumber\quad.\end{eqnarray}
The maximal contraction is obtained by contracting the $s_1-1$ boxes and
the $j$ boxes, which leaves one with a product: \begin{picture}(36,0)(0,0)\multiframe(0,-2)(10,0){1}(35,10){$s_2-s_1$}\end{picture} $\otimes$ \begin{picture}(30,0)(0,0)\multiframe(0,-2)(10,0){1}(30,10){$i-j$}
\end{picture} ,
which always involves a totally symmetric component. Explicitly, this reads:
\begin{eqnarray}
U^{(i)\mu_1\nu_1|...|\mu_j\nu_j|\mu_{j+1}\beta_1|...|\mu_i\beta_{i-j}|\mu_{i+1}...\mu_{s_1-1}}U^{(j)}_{\mu_1\nu_1|...|\mu_j\nu_j|\mu_{j+1}...\mu_{s_2-1}}\quad.
\end{eqnarray}
The $\beta$ indices are free and there are $s_2-s_1$ free $\mu$ indices. The minimal number $N_{min}$ of free indices in this case is thus $s_2-s_1+i-j$.\end{itemize} The two cases can be gathered as \begin{eqnarray}N_{min}=s_2-s_1+|i-j|.\end{eqnarray}
\item[\textbf{B.}]
Secondly, the maximal number of free indices that can be symmetrized in a
product $U^{(i)}U^{(j)}$ may also be studied:
\begin{itemize}
\item[\textbf{B.1.}] If $i\leqslant j<s_1$, then $j$ pairs have to be contracted:\begin{eqnarray}U^{(i)}:\begin{picture}(95,0)(0,0)\multiframe(0,0)(10,0){1}(30,10){$j$}\multiframe(30.5,0)(10,0){1}(60,10){$s_1-j-1$}\multiframe(0,-10.5)(10,0){1}(20,10){$i$}\end{picture},\ U^{(j)}:\begin{picture}(115,0)(0,0)\multiframe(0,0)(10,0){1}(110,10){$s_2-1$}\multiframe(0,-10.5)(10,0){1}(30,10){$j$}\end{picture}\quad.\end{eqnarray}
If one contracts less than $j$ pairs of indices,
some indices remain in the second line of $U^{(j)}$ and the result cannot
contain any totally symmetric component. This leaves us with $s_1+s_2+i-j-2$
free indices.
\item[\textbf{B.2.}] If $i\geqslant j$, in the same way $i$ pairs have to be contracted, leaving $s_1+s_2+j-i-2$ free indices.\end{itemize}
Thus, the maximal number of free and symmetrizable indices in the tensor
product of the two ghost tensors is
\begin{eqnarray}N_{max}=s_1+s_2-|i-j|-2.
\end{eqnarray}
\end{enumerate}
\subsection{Bound on the difference of the number of derivatives}
After these general considerations on allowed products of two ghost tensors $U^{(i)}$ and $U^{(j)}$, let us now consider again a candidate for $a_2$, for a configuration $s-s'-s''$ with $s\leqslant s'\leqslant s''$.
Firstly, it is noticed that there are no nonabelian deformations if $s''\geqslant s+s'$. For example, there is no way of building a $1-1-s$ deformation if $s\geqslant 2$, or a $2-2-s$ deformation with $s\geqslant 4$. This property comes from\footnote{Notice that this property follows from purely diagrammatic reasoning and therefore also applies in $AdS$. Indeed, it can be checked that the structure constants of the higher-spin algebra in \cite{Vasiliev:2003ev} obey to this bound.} the fact that the product of two Young diagrams whose first rows have lengths $s$ and $s'$ cannot contain a Young diagram whose first row has length
length $s''\geqslant s+s'\,$.
Secondly, we can show a stronger property that involves the numbers of derivatives $i$ and $j$. Three cases have to be studied, related to the spin of the antifield. In the case of a spin-$s$ antifield $C^{*\mu_1...\mu_{s-1}}$, the minimum number of free indices in the product $U^{(i)}U^{(j)}$ is $s''-s'+|i-j|$. In order for $a_2$ to be Lorentz-invariant, every index must be contracted, hence the former number must be lower or equal to the number of indices of the antifield. We thus obtain the relation: $s''-s'+|i-j|\leqslant s-1$. In the case of a spin $s'$ antifield, the same argument can be applied, it leads to the relation $s''-s+|i-j|\leqslant s'-1$, which is the same as the first one. Finally, in the case of a spin $s''$ antifield, the minimal condition $s'-s+|i-j|\leqslant s''-1$ is always satisfied, since $i<s$ and $j<s$ imply $|i-j|-s<0$ while, moreover, $s'\leqslant s''$.
On the other hand, in this case, we have to consider the fact that the maximum number of free symmetrizable indices must be greater or equal than the number of indices of the antifield: $s+s'-|i-j|-2\geqslant s''-1$, and we obtain once again the same condition.
Thus for any combination of the fields, the spins have to satisfy the inequality: \begin{eqnarray}s+s'-s''>|i-j|\geqslant0\quad.\end{eqnarray}
This provides an upper bound on the difference between the numbers of derivatives in the two ghost tensors.
\subsection{Conditions on the total number of derivatives}
If we want to build Lorentz-invariant and parity-even expressions, the total number of indices has to be even. For an antifield of spin $s_3$ and ghost tensors $U^{(i)}$ of spin $s_1$ and $U^{(j)}$ of spin $s_2$, the numbers of indices are $s_3-1$, $s_1+i-1$ and $s_2+j-1$, for a total of $s_1+s_2+s_3+i+j-3$. Thus, we find that, for a configuration $s\leqslant s'\leqslant s''\,$:
\begin{eqnarray}
s+s'+s''+i+j\equiv 1(mod\ 2)\quad.
\end{eqnarray}
Furthermore, let us emphasize that the total number of derivatives $\,i+j\,$
is bounded. As was mentioned in Subsection \ref{product}, the numbers $i$ and $j$ must be strictly lower than the spins of the two ghost tensors in order for the free indices to be symmetrizable. If $U^{(i)}$ is of spin $s_1$ and $U^{(j)}$ is of spin $s_2$ with $s_1\leqslant s_2$, then $i\leqslant s_1-1$ and $j\leqslant s_1-1$. Thus, we obtain the condition $i+j\leqslant 2s_1-2\,$.
If we consider a candidate for $a_2\,$, this condition immediately tells that
\begin{eqnarray}
i+j\leqslant 2s'-2\quad.
\end{eqnarray}
More, precisely, if the spin-$s$ antifield is considered,
then the upper bound is $2s'-2\,$.
If either the spin $s'$ or $s''$ antifield is considered, the upper bound
is even lower: $i+j\leqslant 2s-2\,$.
\subsection{A general theorem and a particular candidate}\label{thma2}
Let us summarize all previous considerations in the following theorem:
\newpage
\begin{theorem}Given a cubic configuration of fields with spins $s\leqslant s'\leqslant s''$, the possible Poincar\'e invariants $a_2=C^* U^{(i)}U^{(j)}\,d^nx$ are contractions of an undifferentiated antifield number-$2$ antighost and of two ghost tensors, involving $i$ and $j$ derivatives. The spins and the numbers of derivatives have to satisfy the following properties:\begin{itemize}
\item $0\leqslant |i-j|<s+s'-s''$
\item $s+s'+s''+i+j$ is odd
\item In the case of a spin-$s$ antifield: $i+j\leqslant 2s'-2$\\ In the case of a spin $s'$ or $s''$ antifield: $i+j\leqslant 2s-2$
\end{itemize}
\end{theorem}
To end up this section, let us mention that the candidate $a_2$ with the highest number of derivatives $i+j=2s'-2$ always satisfies
Eq. (\ref{secondorder}) due to the large number of derivatives
involved in it.
Let us also show that the same candidate $a_2$ satisfies
Eq.(\ref{eqagh1}), \textit{i.e.} that a corresponding first-order deformation of the gauge transformations exist. This is less obviously seen than the previous property: We do not provide the corresponding $a_1$ explicitly but the equation ensures that it exists. In the case of an even number of derivatives (in other words when the sum $s+s'+s''$ is odd), the candidate with $2s'-2$ derivatives reads:
\begin{eqnarray}a_2&=&C^{*\mu_1...\mu_{s-1}}U^{(s'-1)}_{\alpha_1\rho_1|...|\alpha_\lambda\rho_\lambda|\mu_1\rho_{\lambda+1}|...|\mu_{s'-\lambda-1}\rho_{s'-1}}\times\nonumber\\&&\times\, U^{(s'-1)\alpha_1\rho_1|...|\alpha_\lambda\rho_\lambda|\phantom{\mu_{s'-\lambda}}\rho_{\lambda+1}|...|\phantom{\mu_{2s'-2\lambda-2}}\rho_{s'-1}|}_{\,\phantom{(s'-1)\alpha_1\rho_1|...|\alpha_\lambda\rho_\lambda|}\mu_{s'-\lambda}\phantom{\rho_{\lambda+1}|...|}\mu_{2s'-2\lambda-2}\phantom{\rho_{s'-1}|}\mu_{2s'-2\lambda-1}...\mu_{s-1}}\,d^n x\quad,\end{eqnarray}
where $\lambda=\frac{s'+s''-s-1}{2}$. In terms of Young diagrams, this contraction can be seen as follows: \begin{eqnarray}
C^*:\begin{picture}(125,0)(0,0)\multiframe(0,0)(10,0){1}(50,10){$s'-\lambda-1$}\multiframe(50.5,0)(10,0){1}(70,10){$s''-\lambda-1$}\end{picture} , U^{(s'-1)}_{s'}:\begin{picture}(75,0)(0,0)\multiframe(0,0)(10,0){1}(20,10){$\lambda$}\multiframe(20.5,0)(10,0){1}(49.5,10){$s'-\lambda-1$}\multiframe(0,-10.5)(10,0){1}(70,10){$s'-1$}\end{picture} , U^{(s'-1)}_{s''}:\begin{picture}(85,0)(0,0)\multiframe(0,0)(10,0){1}(20,10){$\lambda$}\multiframe(20.5,0)(10,0){1}(70,10){$s''-\lambda-1$}\multiframe(0,-10.5)(10,0){1}(70,10){$s'-1$}\end{picture}\quad.\end{eqnarray}
The variation of this expression under delta takes the form:
\begin{eqnarray}
\nonumber
\delta a_2=d(...)+s\Big[\phi^{*\mu_1...\mu_s}
-\frac{(s-1)(s-2)}{2(n+2s-6)}\eta^{(\mu_1\mu_2}
\phi^*{}'{}^{\mu_3...\mu_{s-1})\mu_s}\Big]\partial_{\mu_s}
\Big[U^{(s'-1)}U^{(s'-1)}\Big]\quad.
\end{eqnarray}
The action of $\partial_{\mu_s}$ on the spin $s'$ tensor $U^{(s'-1)}$ is automatically $\gamma$-exact because it is not possible to take one more curl. Actually, the action of $\partial_{\mu_s}$ on the spin $s''$ tensor $U^{(s'-1)}$ is also $\gamma$-exact because the contraction of all free $\mu$ indices with the symmetric indices of the factor linear in the antifield $\phi^*$ prevent any more curl.
The case of an even sum $s+s'+s''$ is a bit more complicated. There are two possible terms, that have to be proportional in order for $a_1$ to exist: \begin{eqnarray}a_2&=&\alpha\, C^{\mu_1...\mu_s}U^{(s'-1)}_{\alpha_1\rho_1|...|\alpha_\lambda\rho_\lambda|\mu_1\rho_{\lambda+1}|...|\mu_{s'-\lambda-1}\rho_{s'-1}}\times\nonumber\\&&\quad\times\, U^{(s'-2)\alpha_1\rho_1|...|\alpha_\lambda\rho_\lambda|\phantom{\mu_{s'-\lambda}}\rho_{\lambda+1}|...|\phantom{\mu_{2s'-2\lambda-3}}\rho_{s'-2}|\rho_{s'-1}}_{\,\phantom{(s'-2)\alpha_1\rho_1|...|\alpha_\lambda\rho_\lambda|}\mu_{s'-\lambda}\phantom{\rho_{\lambda+1}|...|}\mu_{2s'-2\lambda-3}\phantom{\rho_{s'-2}|\rho_{s'-1}}\mu_{2s'-2\lambda-2}...\mu_{s-1}}\,d^n x\ \nonumber\\&&
+\,\beta\, C^{\mu_1...\mu_s}U^{(s'-2)}_{\alpha_1\rho_1|...|\alpha_\lambda\rho_\lambda|\mu_1\rho_{\lambda+1}|...|\mu_{s'-\lambda-2}\rho_{s'-2}|\rho_{s'-1}}\times\nonumber\\&&\ \,\quad\times\, U^{(s'-1)\alpha_1\rho_1|...|\alpha_\lambda\rho_\lambda|\phantom{\mu_{s'-\lambda}-1}\rho_{\lambda+1}|...|\phantom{\mu_{2s'-2\lambda-2}}\rho_{s'-1}|}_{\,\phantom{(s'-1)\alpha_1\rho_1|...|\alpha_\lambda\rho_\lambda|}\mu_{s'-\lambda-1}\phantom{\rho_{\lambda+1}|...|}\mu_{2s'-2\lambda-3}\phantom{\rho_{s'-1}|}\mu_{2s'-2\lambda-2}...\mu_{s-1}}\,d^n x\quad,\end{eqnarray} where $\lambda=(s'+s''-s-2)/2$
and $\alpha,\beta$ are coefficients. In terms of Young diagrams, these contractions read:\begin{eqnarray}&&
C^*:\begin{picture}(125,0)(0,0)\multiframe(0,0)(10,0){1}(50,10){$s'-\lambda-1$}\multiframe(50.5,0)(10,0){1}(70,10){$s''-\lambda-2$}\end{picture} , U^{(s'-1)}_{s'}:\begin{picture}(75,0)(0,0)\multiframe(0,0)(10,0){1}(20,10){$\lambda$}\multiframe(20.5,0)(10,0){1}(50,10){$s'-\lambda-1$}\multiframe(0,-10.5)(10,0){1}(60,10){$s'-2$}\multiframe(60.5,-10.5)(10,0){1}(10,10){1}\end{picture} , U^{(s'-2)}_{s''}:\begin{picture}(100,0)(0,0)\multiframe(0,0)(10,0){1}(20,10){$\lambda$}\multiframe(20.5,0)(10,0){1}(70,10){$s''-\lambda-2$}\multiframe(91,0)(10,0){1}(10,10){1}\multiframe(0,-10.5)(10,0){1}(60,10){$s'-2$}\end{picture}\nonumber\quad,\\ \textrm{and}&&\nonumber\\&& C^*: \begin{picture}(125,0)(0,0)\multiframe(0,0)(10,0){1}(50,10){$s'-\lambda-2$}\multiframe(50.5,0)(10,0){1}(70,10){$s''-\lambda-1$}\end{picture} , U^{(s'-2)}_{s'}:\begin{picture}(85,0)(0,0)\multiframe(0,-10.5)(10,0){1}(20,10){$\lambda$}\multiframe(20.5,-10.5)(10,0){1}(49.5,10){$s'-\lambda-2$}\multiframe(0,0)(10,0){1}(80,10){$s'-1$}\end{picture} , U^{(s'-1)}_{s''}:\begin{picture}(90,0)(0,0)\multiframe(0,0)(10,0){1}(20,10){$\lambda$}\multiframe(20.5,0)(10,0){1}(70,10){$s''-\lambda-1$}\multiframe(0,-10.5)(10,0){1}(80,10){$s'-1$}\end{picture}\quad.\nonumber
\end{eqnarray}
This time, the computation of $\delta a_2$ consists of four terms. The term involving the derivative of the spin-$s'$ tensor $U_{s'}^{(s'-1)}$ is automatically $\gamma$-exact, and the term where the spin-$s''$ tensor $U_{s''}^{(s'-1)}$ is differentiated is $\gamma$-exact, thanks to the same arguments as for the odd case. On the other hand, the terms where the $U^{(s'-2)}$ tensors are differentiated are problematic. Fortunately, the non-$\gamma$-exact terms that appear are the same in the two expressions and the coefficients $\alpha$ and $\beta$ can be fitted to obtain a $\gamma$-exact result.
\section{Proof of the strong obstruction on the BBvD vertex}\label{sec:obstr}
The BBvD first order deformation \cite{Berends:1984wp} has been obtained in the antifield formulation in \cite{Bekaert:2006jf}. We denote the spin-3 ghost tensors $T^A_{\mu\nu|\rho}:=U^{(1)A}_{\mu\nu|\rho}$ and $U^A_{\mu\nu|\rho\sigma}:=U^{(2)A}_{\mu\nu|\rho\sigma}$. The antifield number $\ 2$ component $a_2$ of the BBvD deformation contains $2$ derivatives and reads:
\begin{eqnarray}a_{2,BBvD}=f^A_{\phantom{A}BC}C^{*\mu\nu}_A\left[T^B_{\mu\alpha|\beta}T^{C\,\alpha|\beta}_\nu-2\,T^B_{\mu\alpha|\beta}T^{C\,\beta|\alpha}_\nu+\frac{3}{2}\,C^{B\alpha\beta}U^C_{\mu\alpha|\nu\beta}\right]d^n x+\gamma c_2\quad,\end{eqnarray} where the capital internal indices span the multiplet of spin-3 fields.
The corresponding cubic vertex $a_{0,BBvD}$ contains $3$ derivatives.
It has been showed \cite{Bekaert:2006jf} that the second order expression, $(a_{2,BBvD},a_{2,BBvD})$ as in Equ.(\ref{secondorder}), presents an obstruction containing terms of the structure $C^*TTU$ and $C^*CUU$, that are not $\gamma$-exact modulo $d$, and cannot be eliminated. The coefficient of the obstruction is $f_{ABC} f^A_{DE}$, whose vanishing implies the vanishing of the deformation itself. Let us notice that, in dimension 3, since $U^A_{\alpha\beta|\gamma\delta}\equiv0$, the BBvD candidate passes the test. In dimension 4, some Schouten identities could imply the weaker associativity condition $f^{}_{AB[C}f^A_{D]E}=0$, however, this still implies the vanishing of $f_{ABC}$ in the end.
Furthermore, a new nonabelian cubic vertex with 5 derivatives was found in \cite{Bekaert:2006jf}. However, it vanishes in dimension 3 where the BBvD candidate therefore involves the maximal possible number of derivatives. In dimension 4, it has been showed that Schouten identities imply the vanishing of the corresponding component of $a_2$, thus the 5-derivative deformation is Abelian in that case.
Let us remark that the case of dimension 3 is a bit special, since traceless tensors associated with a Young diagram whose first two columns have length 2 identically vanish.\footnote{More generally, in dimension $n$, any tensor associated with an irreducible representation of $\mathfrak{o}(n)$ (and thus traceless), whose Young diagram is such that the sum of the heights of the first two columns is greater than $n$, identically vanish (see \cite{Hamermesch}, page 394). Let us remark that, for $n\geqslant 4$, two-row tensors, such as the traceless part of the curvature or the strictly non-$\gamma$-exact ghost tensors, are never constrained by this condition.} This implies that every fieldstrength vanishes on-shell, which only allows topological theories, for spin $\geqslant 2\,$.
We now want to prove that no other $a_2$'s can provide the same kind of terms that could compensate the obstruction. First, the antibracket $(a_{2,BBvD},a_{2,BBvD})$ is of course quartic in the spin-3 fields (in the extended meaning of fields, ghosts or antifields) and it contains four derivatives. It can only be compensated by terms in another antibracket $(a_2,a_2)$ which have exactly the same structure. The only possibility of getting terms quartic in the spin-3 fields is to take the $(a_2,a_2)$ of two expressions with the same spin configuration $s-3-3$. Then, the first rule of our general theorem ensures that $1\leqslant s \leqslant 5$. The nonabelian $1-3-3$ and $2-3-3$ deformations have been completely classified \cite{Boulanger:2006gr,Boulanger:2008tg}: in both cases, there is only one solution, whose $a_2$ is linear in the antifield with lowest spin (1 or 2) and antifield number $2$. These candidates satisfy trivially $(a_2,a_2)=0$, hence they cannot help for the BBvD obstruction.
To complete the argument, we have to investigate the $3-3-4$ and $3-3-5$ cases. It is rather simple: we will prove that the only $a_2$ candidates that are related to an $a_1$ contain
at least three derivatives, which is sufficient to be sure that no obstruction containing four derivatives will arise. The results about those two cases are presented in the next two subsections. The results are sufficient to establish the inconsistency of the BBvD deformation in dimension greater than three, in the parity-invariant case, thereby invalidating the hopes expressed by the authors of \cite{Berends:1984wp,Berends:1984rq}
concerning a possible solution of their problem by the addition of
totally symmetric higher spin contributions. Therefore, in flat spacetime, their spin-3 self coupling is definitely inconsistent and no
totally symmetric higher spin field can cure this problem contrary to the general belief. It is only in $(A)dS$ that this candidate can play a role, as suggested by the Fradkin-Vasiliev cubic vertices \cite{Fradkin:1987ks,Fradkin:1987qy}.
\vspace{1mm}
\noindent{\bf{Remark:}} The $a_2$ components that are considered in the sequel for the $3-3-4$ and $3-3-5$ cases are not proved to be part of consistent first order solutions. Anyway, since we seek a negative result, it is obvious that, if the obstruction remains when considering all of the candidates, it would remain \textit{a fortiori} if these candidates are obstructed at first order.
\subsection{Study of $a_2$ in the $3-3-4$ case}
Let us use the theorem of Subsection \ref{thma2}, with $s=s'=3$ and $s''=4$. The sum of the spins is even, thus the number of derivatives in $a_2$ has to be odd. The maximum is $2s'-3=3$. Furthermore, the difference between the numbers of derivatives acting on the two ghosts obeys $|i-j|<s+s'-s''=2$, and is thus equal to 1. The possible strictly non-$\gamma$-exact Lorentz-invariant expressions with one derivative read: \begin{eqnarray}\stackrel{(1)}{t}{}^{\!\!AB}=C^{*\mu\nu\rho}C^{A\alpha}_{\mu}T^B_{\alpha\nu|\rho}\ ,\ \stackrel{(2)}{t}{}^{\!\!AB}=C^{*A\mu\nu} T^B_{\mu\alpha|\beta} C_\nu^{\ \alpha\beta}\ ,\ \stackrel{(3)}{t}{}^{\!\!AB}=C^{*A\mu\nu} C^{B\rho\sigma} U^{(1)}_{\mu\rho|\nu\sigma}\quad.\end{eqnarray} Those with three derivatives are: \begin{eqnarray}\stackrel{(4)}{t}{}^{\!\!AB}=C^{*\mu\nu\rho}T_{\mu}^{A\alpha|\beta}U_{\nu\alpha|\rho\beta}^B\ ,\ \stackrel{(5)}{t}{}^{\!\!AB}=C^{*A\mu\nu}T^{B\alpha\beta|\gamma}U^{(2)}_{\alpha\beta|\gamma\mu|\nu}\ ,\ \stackrel{(6)}{t}{}^{\!\!AB}=C^{*A\mu\nu}U^B_{\alpha\beta|\gamma\mu}U^{(1)\alpha\beta|\gamma}_{\phantom{(1)\alpha\beta|\gamma}\nu}\quad.\end{eqnarray}The spin-4 internal indices have not been written explicitly since no symmetries can arise involving them (similarly, in the next section, the spin-5 indices are not written as well).
Let us check that the candidates with three derivatives are related to an $a_1$: \begin{eqnarray}\delta\stackrel{(4)}{t}{}^{\!\!AB}=\mbox{divergence}+\gamma(...)+4\phi^{*\mu\nu\rho\sigma}U^{A\alpha|\phantom{\sigma}\beta}_{\mu\phantom{\alpha|}\sigma}U^B_{\nu\alpha|\rho\beta}-\frac{6}{n+2}\phi^*{}'{}^{\rho\sigma}U^{A\nu\alpha|\phantom{\sigma}\beta}_{\phantom{A\nu\alpha|}\sigma}U^B_{\nu\alpha|\rho\beta}\quad.\end{eqnarray} This term is antisymmetric in $AB$, thus a symmetric set of coefficients ensures the vanishing of the non-$\gamma$-exact terms. The variation under $\delta$ of the two other terms provides the same non-$\gamma$-exact term $\phi^{*A\mu\nu\rho}U^{B\alpha\beta|\phantom{\rho}\gamma}_{\phantom{B\alpha\beta|}\rho}U^{(2)}_{\alpha\beta|\gamma\mu|\nu}$, so they vanish if $\stackrel{(5)}{t}{}^{\!\!AB}$ and $\stackrel{(6)}{t}{}^{\!\!AB}$ have opposite coefficients. Finally, we get as candidates with three derivatives: \begin{eqnarray}a_{2,3}=k_{(AB)}C^{*\mu\nu\rho}T_{\mu}^{A\alpha|\beta}U_{\nu\alpha|\rho\beta}^B d^n x+l_{AB}C^{*A\mu\nu}\Big[T^{B\alpha\beta|\gamma}U^{(2)}_{\alpha\beta|\gamma\mu|\nu}-U^B_{\alpha\beta|\gamma\mu}U^{(1)\alpha\beta|\gamma}_{\phantom{(1)\alpha\beta|\gamma}\nu}\Big]d^n x\quad.\end{eqnarray} On the other hand, the candidates involving one derivative are obstructed: \begin{eqnarray}\delta \stackrel{(1)}{t}{}^{\!\!AB}=\mbox{divergence}+\gamma(...)+4\Big(\phi^{*\mu\nu\rho\sigma}-\frac{3}{n+2}\eta^{(\mu\nu}\phi^*{}'{}^{\rho)\sigma}\Big)\Big[-T^{A\alpha}_{\phantom{A\alpha}\sigma|\mu}T^B_{\alpha\nu|\rho}+C^{A\alpha}_{\mu}U^B_{\alpha\nu|\sigma\rho}\Big]\quad.\end{eqnarray} All the terms vanish if $\stackrel{(1)}{t}{}^{\!\!AB}$ is multiplied by a symmetric coefficient, except one proportional to the trace of $\phi^*$: $\frac{-4}{n+2}\phi^*{}'{}^{\nu\sigma}C^{A\alpha\rho}U^{B}_{\alpha\nu|\sigma\rho}$. This obstruction cannot be removed. The variation under $\delta$ of $\stackrel{(2)}{t}{}^{\!\!AB}$ and $\stackrel{(3)}{t}{}^{\!\!AB}$ contains the obstructions $\phi^{*A\mu\nu\rho}U^B_{\mu\alpha|\rho\beta}C_\nu^{\phantom{\nu}\alpha\beta}$ and $\phi^{*A\mu\nu\rho}C^{\alpha\beta}U^{(2)}_{\mu\alpha|\nu\beta|\rho}$. Finally, the only possible $3-3-4$ deformation contains three derivatives in $a_2$. Even if the vertex exists, which is not sure, the only terms in $(a_{2,3},a_{2,3})$ contain six derivatives. This cannot remove the obstruction of the BBvD deformation.
\subsection{Study of $a_2$ in the $3-3-5$ case}
The theorem of Subsection \ref{thma2} ensures that the number of derivatives in $a_2$ is even, and is not greater than 4. Furthermore the two ghosts bear the same number of derivatives, since $|i-j|<3+3-5=1$. There are candidates with four, two and zero derivatives. Once again, only the candidates with four derivatives satisfy Eq.(\ref{eqagh1}). The possible terms with no derivatives are: \begin{eqnarray} \stackrel{(1)}{u}{}^{\!\!AB}=C^{*\mu\nu\rho\sigma}C^A_{\mu\nu}C^B_{\rho\sigma}\quad{\textrm{and}}\quad \stackrel{(2)}{u}{}^{\!\!AB}=C^{*A\mu\nu}C^{B\rho\sigma}C_{\mu\nu\rho\sigma}\quad.\end{eqnarray}
Those with two derivatives are: \begin{eqnarray}\stackrel{(3)}{u}{}^{\!\!AB}=C^{*\mu\nu\rho\sigma}T^{A\alpha}_{\phantom{A\alpha}\mu|\nu}T^B_{\alpha\rho|\sigma}\quad{\textrm{and}}\quad \stackrel{(4)}{u}{}^{\!\!AB}=C^{*A\mu\nu}T^{B\alpha\beta|\gamma}U^{(1)}_{\alpha\beta|\gamma\mu\nu}\quad.\end{eqnarray} Those with four derivatives are: \begin{eqnarray} \stackrel{(5)}{u}{}^{\!\!AB}=C^{*\mu\nu\rho\sigma}U^{A\alpha|\phantom{\nu}\beta}_{\mu\phantom{\alpha|}\nu}U^B_{\rho\alpha|\sigma\beta}\quad{\textrm{and}}\quad\stackrel{(6)}{u}{}^{\!\!AB}=C^{*\mu\nu}U^{\alpha\beta|\gamma\delta}U^{(2)}_{\alpha\beta|\gamma\delta|\mu\nu}\quad.\end{eqnarray}Let us notice that $\stackrel{(1)}{u}{}^{\!\!AB}$, $\stackrel{(3)}{u}{}^{\!\!AB}$ and $\stackrel{(5)}{u}{}^{\!\!AB}$ are naturally antisymmetric over $AB$. It is quite obvious that $\delta\stackrel{(5)}{u}{}^{\!\!AB}$ is $\gamma$-exact modulo $d$ because the third derivative of the spin-3 ghost are $\gamma$-exact. Then, we can consider $\delta\stackrel{(6)}{u}{}^{\!\!AB}$, which is $\gamma$-exact modulo $d$, for the same reason than the previous one and because $\partial_{(\rho}U^{(2)\alpha\beta|\gamma\delta|}_{\phantom{(2)\alpha\beta|\gamma\delta|}\mu\nu)}$ is $\gamma$-exact. On the other hand, obstructions arise for any of the other candidates. For $\stackrel{(3)}{u}{}^{\!\!AB}$, one of the trace term remains, which is proportional to $\phi^*{}'{}^{\mu\nu\tau}U^A_{\alpha\mu|\tau\sigma}T^{B\alpha\ |\sigma}_{\phantom{B\alpha}\nu}$. For $\stackrel{(4)}{u}{}^{\!\!AB}$, the obstruction consists of two terms, proportional to $\phi^{*A\mu\nu\rho}T^{B\alpha\beta|\gamma}U^{(2)}_{\alpha\beta|\gamma\mu|\nu\rho}$ and $C^{*A\mu\nu\rho}U^{B\alpha\beta|\phantom{\rho}\gamma}_{\phantom{B\alpha\beta|}\rho}U^{(1)}_{\alpha\beta|\gamma\mu\nu}$. Finally, with no derivatives, the obstruction of $\stackrel{(1)}{u}{}^{\!\!AB}$ arises once again in the trace terms, it is proportional to $\phi^*{}'{}^{\mu\nu\rho}T^A_{\alpha\mu|\nu}C^{B\alpha}_{\phantom{B\alpha}\rho}$. The obstruction of $\stackrel{(2)}{u}{}^{\!\!AB}$ consists of two terms proportional to $\phi^{*A\mu\nu\rho}C^{B\alpha\beta}U^{(1)}_{\rho\alpha|\beta\mu\nu}$ and $\phi^{*A\mu\nu\rho}T^{B\alpha|\beta}_{\rho}C_{\mu\nu\alpha\beta}$. None of those obstructions can be removed, the only possible cubic $a_2$ thus involves four derivatives. Thus, any $(a_2,a_2)$ term involves eight derivatives, this can of course not remove the BBvD obstruction. Since we have considered a spin greater than four, we are not sure if the cubic deformations are the only possible ones, but any solution of degree higher than three will provide terms of power higher than four in $(a_2,a_2)$, which can not compensate the BBvD obstruction either.
\section{Conclusion}\label{concl}
Within the antifield formalism and in the case of cubic vertices between symmetric tensor gauge fields of any integer spins, we have introduced a set of criteria for the construction of consistent
deformations of the gauge algebra.
Equivalently, these criteria are conditions on the structure constants of the
gauge algebra at first order in the coupling constants. We have then showed
that the Berends--Burgers--van Dam spin-3 vertex is obstructed at second order
in the coupling constants, even if one introduces other symmetric tensor gauge
fields in the theory. This invalidates, in Minkowski spacetime, the argument
according to which the obstructions arising for a given set of values of the
spins can be cured by terms involving higher values. This argument is related
to the standard lore that an infinite tower of fields with unbounded spin is
needed in any consistent higher-spin gauge theory. While this general
expectation is not questioned, our result
confirms some doubts about the mere existence of any consistent nonabelian
Lagrangian formulation for higher-spin
gauge fields in four-dimensional\footnote{The pure spin-$3$ non Abelian cubic
vertex found in
\cite{Bekaert:2006jf}
exists only in higher dimensions ($n>4$) but this one is \textit{not}
obstructed at the level of the gauge algebra. Nevertheless, the existence of a
corresponding quartic vertex remains an open issue.} Minkowski spacetime, which
would be obtained as a perturbative local deformation of Fronsdal's theory.
For flat spacetimes of higher dimensions, our results suggest that nonabelian cubic vertices containing only totally symmetric gauge fields and involving a number of derivatives which does \emph{not} saturate the upper bound that we found, would be inconsistent.
More precisely, our argument essentially
relies on the numbers of derivatives. Consistent first order vertices must involve a minimal number of derivatives. This minimum number increases with the values of the spin of the three fields contained in the cubic vertex. The number of derivatives is a good grading in flat spacetime, so the second-order equations involving different types of vertices are most of the times linearly independent (because they contain different numbers of derivatives). We can conjecture that many consistent cubic deformations in Minkowski spacetime are strongly obstructed in the same way.
However, in $(A)dS$ the number of derivatives is not a proper grading
and we expect that the obstructions exhibited for the flat-spacetime vertices
do not show up in $(A)dS\,$, so that the Fradkin--Vasiliev cubic Lagrangian
\cite{Fradkin:1987ks,Fradkin:1987qy} (see also \cite{Vasiliev:2001wa,Alkalaev:2002rq})
could be completed to give a fully consistent nonabelian Lagrangian theory to all orders in the coupling constant.
\section*{Acknowledgments}
We thank G. Barnich, M. Henneaux, A. Sagnotti, Ph. Spindel and P. Sundell
for discussions.
N.B. is a Research Associate of the Fonds de la Recherche Scientifique--FNRS, Belgium.
\providecommand{\href}[2]{#2}\begingroup\raggedright |
\section{Introduction}\label{intro}
30~Doradus in the Large Magellanic Cloud (LMC) is the largest
star-forming region in the Local Group. As an archetypal,
small-scale `starburst', 30~Dor provides us with an excellent laboratory
in which to study star formation and stellar evolution, while also giving us
potential insights into the nature of distant super-star-clusters for
which we only have integrated properties. Extensive ground-based
optical imaging and spectroscopy has been used to study the initial
mass function (IMF), reddening, star-formation history, stellar
content and kinematics in 30~Dor \citep[e.g.][]{m85,p93,pg93,wb97,s98,b01}.
Meanwhile, near-IR observations with the {\em Hubble Space Telescope
(HST)} have provided evidence of triggered star-formation, showing the region to
be a two-stage starburst \citep{wal99,wal02}.
At the centre of 30~Dor is the dense star cluster R136, with stellar
ages for the most massive stars in the range of 1-2\,Myr \citep{mh98} and a total stellar mass in
the range of $\sim$0.35-1$\times10^{5}$\,M$_\odot$
\citep[e.g.][]{mg03,ng07,a09}, depending on the low-mass form of the mass function,
putting it on a par with some of the clusters found in starburst and interacting galaxies such as M51, M82
and the Antennae. However, the core of R136 is too dense for
traditional (seeing-limited) ground-based techniques.
Only with the arrival of {\em HST} was R136 resolved in optical and UV
images \citep{c92,dm93,h95,h97,s00}, with follow-up spectroscopy revealing a
hitherto unprecedented concentration of the earliest O-type stars
\citep{mh98}. Star counts from the optical {\em HST} images revealed
that the luminosity profile of R136 appears to be best described by
two components, with a break at 10$''$ \citep{mg03}. New results from
{\em F160W} imaging (equivalent to {$H$}-band) with the {\em HST} Near
Infrared Camera and Multi-Object Spectrometer (NICMOS) fit the profile
with a single component \citep{a09}, but only in
the inner 2\,pc (8\farcs25). Whether the second component seen in the
optical data is a manifestation of the `excess light' predicted to
originate from rapid gas removal in the early stages of cluster
evolution \citep{bg06} remains an open question as there is significant and
variable extinction across the cluster.
As part of the technology development plan towards the European
Extremely Large Telescope (E-ELT), the Multi-conjugate Adaptive optics
Demonstrator \citep[MAD,][]{em07} was developed as a visiting
instrument for the Very Large Telescope (VLT). The offer of Science
Demonstration (SD) observations with MAD presented the perfect
opportunity to obtain imaging of R136 at unprecedented angular
resolution at near-IR wavelengths. In particular, MAD has the power
to penetrate the gas and dust more successfully than in the optical
{\em HST} images, at comparable angular resolution, to provide
empirical constraints on the outer component on the luminosity
profile. Determining if R136 is an expanding group or a
dynamically-stable star cluster would serve as an important ingredient
in the debate on the importance of `infant mortality' of young
clusters \citep[e.g.][]{ll03,bg05,fcw05,gb06,bk07,bggt08}.
Here we present
the MAD SD observations, which deliver a cleaner point spread function
(PSF) than NICMOS, at finer angular resolution, and over a larger total field. The instrumental
performance of MAD is discussed in Section~\ref{obsdata}, with the
astrometric and photometric calibration of the data detailed in
Sections~\ref{astrometry} and \ref{photometry}, respectively.
Following discussion of seven of the candidate young stellar objects (YSOs)
reported by \citet[][hereafter GC09]{g09} that lie within the MAD
fields (Section~\ref{ysos}), we then investigate the radial luminosity
profile of R136 via a combination of star counts and integrated-light
measurements (Section~\ref{profiles}), discussing its implications in
the context of cluster formation and observations of distant
unresolved clusters.
\section{Observations \& Data Reduction}\label{obsdata}
MAD employs three Shack-Hartmann wavefront sensors to observe three
natural guide stars (NGS) across a 2$'$ circular field, thereby
allowing tomography of the atmospheric turbulence. The turbulence is
then corrected using two deformable mirrors (operating at $\sim$400
Hz), one conjugated to the ground-layer (i.e. 0\,km), the second
conjugated to 8.5\,km above the telescope.
The high resolution, near-IR camera used with MAD is the CAmera for
Multi Conjugate Adaptive Optics \citep[CAMCAO,][]{camcao}, which
operates over the {\em J}, {$H$}, and $K_{\rm s}$\/ bands, with critical (2
pixel) sampling of the diffraction-limited PSF
at 2.2$\mu$m. The detector is a 2k$\times$2k Hawaii-2 HgCdTe array,
with a pixel scale of 0\farcs028/pixel, giving a field-of-view of 57$''$x57$''$.
A useful feature is that the camera can be moved within the 2$'$ field
without requiring positional offsets of the telescope, meaning that
the adaptive optics (AO) loop can remain closed.
\begin{figure*}
\begin{center}
\includegraphics[scale=0.6]{fig1.pdf}
\caption{$V$-band image of the central part of 30~Dor (approx 3\farcm0
by 2\farcm5) from the Wide-Field Imager (WFI) on the ESO/Max Planck Gesellschaft 2.2-m telescope (from observations
by J.~Alves, B.~Vandame \& Y.~Beletsky). North is towards the top of
the figure, east towards the left. The reference stars used for the
MCAO correction are shown (GS1, 2 \& 3), together with the spatial
extent of the dithered {$H$}-band observations of the three fields.}\label{obs}
\includegraphics[scale=0.55]{fig2.pdf}\\
\caption{MAD $K_{\rm s}$-band image of Field~3. Compared to the overlay on
Figure~\ref{obs} the dithered regions are cropped, with a final field-of-view
of $\sim$52$''$\,$\times$\,46$''$.}\label{kp3}
\end{center}
\end{figure*}
The MAD data presented here were obtained from one of twelve SD
programmes, observed in November 2007 and January
2008. {$H$}- and $K_{\rm s}$-band images were obtained for three pointings in
the inner region of 30~Dor, as shown in Figure~\ref{obs}. The
central co-ordinates for the CAMCAO observations of Field 1 were
$\alpha$\,$=$\,05$^{\rm h}$38$^{\rm m}$46.5$^{\rm s}$, $\delta$\,$=$\,$-$69$^\circ$05$'$52$''$ (J2000.0). Fields 2 and 3 were offset from
this first pointing by approximately $-$25$''$ in right ascension and $\pm$25$''$ in
declination. The three NGS used for wavefront sensing are also shown
in Figure~\ref{obs}; these are Parker \#952 (GS1), \#499 (GS2), and
\#1788 (GS3) with {\em V}\,=\,12.0, 11.9, and 12.0, respectively \citep{p93}.
The combined $K_{\rm s}$-band image of Field~3 is shown in Figure~\ref{kp3}.
The observations are summarised in Table~\ref{obsinfo}. The detector
integration time (DIT) for all of the observations was 2\,s, with 30
integrations (NDIT) for each exposure. Batches of three ({$H$}) and six
($K_{\rm s}$) object and sky frames were interleaved in an A-B-A-B-A-B-A
pattern, yielding total exposures of 12\,min for each field in the
{$H$}\/ band, and 24\,min in $K_{\rm s}$. Each science exposure within each
batch was dithered by 5$''$ -- although this reduces the
effective area of the final combined images, it was intended to
minimise the impact of bad pixels and cosmetic features from the
array. Given the vast spatial extent of 30~Dor, the sky offsets were
somewhat large ($+$12s of right ascension, $+$13$'$ in declination) to
ensure they were uncontaminated by nebulosity. Observations were
halted mid-way through the $K_{\rm s}$\/ exposures for Field 1 on 2008 January 7
due to bad weather, but were completed the following night. Note that
the LMC never rises above an altitude of approx. 45$^\circ$ as viewed
from Paranal, i.e. the {\em minimum} zenith distance of the
observations was $\sim$45$^\circ$, providing a good test of the MAD
performances at moderately large airmass. Indeed, the airmass of the
observations ranged from 1.4 to 1.6. The range of seeing values for
each field, as measured by the Differential Image Motion Monitor
(DIMM) at Paranal, are given in Table~\ref{obsinfo}.
\begin{table*}
\begin{center}
\caption{Summary of the VLT-MAD observations in 30~Doradus. The total exposure times quoted are
for the final combined images.}\label{obsinfo}
\begin{tabular}{lclcccc}
\hline
Pointing & Band & Date & Total Exp. & DIMM range & Image FWHM & $<$FWHM$>$ \\
& & & [min] & [$''$] & [$''$] & [$''$] \\
\hline
Field 1 & $K_{\rm s}$ & 2008/01/07 \& 08 & 22 & 0.4-1.8 & 0.10-0.13 & 0.11 \\
Field 2 & $K_{\rm s}$ & 2008/01/07 & 24 & 0.5-1.1 & 0.08-0.10 & 0.09 \\
Field 3 & $K_{\rm s}$ & 2007/11/27 & 23 & 0.6-1.0 & 0.10-0.20 & 0.14 \\
\hline
Field 1 & {$H$} & 2008/01/08 & 12 & 0.3-0.6 & 0.10-0.12 & 0.11 \\
Field 2 & {$H$} & 2008/01/08 & 12 & 0.9-1.1 & 0.08-0.11 & 0.09 \\
Field 3 & {$H$} & 2008/01/08 & 11 & 0.6-1.6 & 0.08-0.15 & 0.12 \\
\hline
\end{tabular}
\end{center}
\end{table*}
The MAD data were reduced with standard {\sc iraf} routines, using
calibration frames from the SD runs to correct for the dark current,
to flat-field all of the object and sky exposures, and to reject bad
pixels and cosmic rays. Median sky frames were created for each batch
of science observations using the sky frames observed immediately
before and/or after, although (in general) the sky background did not
vary strongly over each sequence of observations. The sky-subtracted
frames were then aligned with each other and combined. At this stage
we omitted one or two images with significantly poorer image quality
in some fields, hence the final exposure times in
Table~\ref{obsinfo}. Note that no objects were saturated in any of the
individual science frames.
\subsection{Image Quality \& Performance Analysis}
Moderately bright stars were used to investigate the image quality
in the three fields. The range and mean of the full-width
half maxima (FWHM) obtained from 32 stars evenly distributed across
each co-added image is summarised in Table~\ref{obsinfo}.
Maps of the Strehl ratio in the $K_{\rm s}$-band images
are shown in Figure~\ref{strehl}. More detailed Strehl maps
(showing the relative positions of the NGS) and FWHM maps in both
photometric bands were presented by \citet{cspie}.
The best Strehl ratio achieved in the $K_{\rm s}$-band (25-30\%) was in the regions closest to
the NGS (as one would expect) in Fields~2 and 3. The performance in
Field~2 is particularly good, with an average FWHM of 0\farcs09 in
both the {$H$}\/ and $K_{\rm s}$\/ images (compared to diffraction limits
[$\lambda$/D] of approx 0\farcs04 and 0\farcs06, respectively).
Although the Strehl is lower in Field~1, this was in the best position
with respect to all three NGS and the correction is very uniform
across the combined $\sim$50$''$x50$''$ image. This is in strong
contrast to Field 3 in which the performance is very good in one
corner, but then steeply declines away from the NGS -- more in keeping
with `classical' AO observations. In general, the performances are
comparable with those found from other MAD observations
\citep[e.g.][]{em07,bouy08}.
\begin{figure*}
\begin{center}
\includegraphics[height=5.3cm]{fig3a.pdf}\hspace{0.1cm}\includegraphics[height=5.3cm]{fig3b.pdf}\hspace{0.1cm}\includegraphics[height=5.3cm]{fig3c.pdf}
\caption{$K_{\rm s}$-band Strehl (\%) maps for Fields~1, 2 and 3 (left to right, respectively). North is at the top, with east at the left.}\label{strehl}
\end{center}
\end{figure*}
\section{ Astrometric Calibration}\label{astrometry}
Astrometric calibration of each field was undertaken using the
\citet{s98} catalogue, recently recalibrated by Brian Skiff\footnote{
http://cdsarc.u-strasbg.fr/viz-bin/Cat?J/A+A/341/9}. These positions
are dependent on the precision of the 2MASS/UCAC2 positions, combined
with the plate-scale/seeing of the original New Technology Telescope
(NTT) observations. The quoted precision on the new Skiff astrometry
is $\sim$0\farcs1 (cf. the 0\farcs028/pixel delivered by MAD). Visual
matches of stars between the Skiff catalogue and the MAD fields were
used to define $\sim$40 well distributed astrometric standards in
Fields~1 and 2. The astrometric calibration of Field~3, in which
there are greater PSF variations, was achieved using 60
visually-matched stars.
\section{Photometric calibration}\label{photometry}
Photometry was performed using standard {\sc daophot} routines in {\sc
iraf} on the combined images. Due to the variations in the PSF
delivered by the AO system across the field, PSF-fitting photometry
was used, with the model PSF ({\sc penny}{\small 2}) allowed to vary
quadratically across the fields. Roughly 30 bright, isolated
stars were used to create a model PSF for each combined image, with
neighbouring stars subtracted if they were contaminating the selected
star. A fitting radius of (2$\times$FWHM)$-$1 pixels
was found to provide the best fits.
Objects with 1.5 $<$ sharpness $< -$1.5, or with $\chi <$ 4
(where $\chi$ is the root-mean-square of the residuals that remain)
were rejected from the final catalogues to reject non-stellar objects,
residual cosmic rays etc., as were objects with instrumental magnitude
errors greater than 0.25$^{\rm m}$.
\subsection{Zero-point Calibrations}
\subsubsection{MAD vs. 2MASS}
Sources from the Two Micron All Sky Survey \citep[2MASS,][]{2mass}
were identified in the MAD frames for photometric calibration.
However, at the excellent angular resolution delivered by MAD many of
the 2MASS `stars' are resolved into asymmetric sources or multiple
components \citep[e.g. Fig.~3,][]{mob08} . As such, only isolated,
apparently single stars with 2MASS photometric qualities of either
`A', `B', or `C' in the relevant band were considered for calibration.
This limited the number of 2MASS stars in each MAD field to eight or
fewer for calibration.
Photometric zero points (ZPs) were determined (principally using
`A'-rated 2MASS stars), with stars at the bright/faint ends of the
MAD images omitted to exclude strongly-exposed stars in the MAD
frames and poorly-exposed stars in 2MASS. The resulting ZPs
are summarised in Table~\ref{zero_points}; we are somewhat
at the mercy of small number statistics, not to mention the issues
of stellar density and complex nebular emission in and around R136.
\begin{table*}
\begin{center}
\caption{Photometric zero-points obtained from the combined image method.
Results are given for the $K_{\rm s}$-band images using 2MASS sources within
the MAD frames, and using the HAWK-I commissioning data, in which the result
for Field~2 was bootstrapped using overlapping stars with Field~1. The {$H$}-band
results were obtained using 2MASS sources for Fields~1 and 3, with Field~2 again
calibrated using overlapping stars with Field~1.}\label{zero_points}
\begin{tabular}{llccc}
\hline
Band & Calibration & Field 1 & Field 2 & Field 3 \\
\hline
$K_{\rm s}$ & 2MASS & 26.78 $\pm$ 0.13 & 26.61 $\pm$ 0.45 & 26.95 $\pm$ 0.25 \\
$K_{\rm s}$ & HAWK-I & 26.69 $\pm$ 0.08 & 27.04 $\pm$ 0.09 & 27.07 $\pm$ 0.11 \\
\hline
{$H$} & 2MASS & 27.09 $\pm$ 0.14 & 27.27 $\pm$ 0.15 & 26.93 $\pm$ 0.08 \\
\hline
\end{tabular}
\end{center}
\end{table*}
\subsubsection {MAD vs. HAWK-I}\label{hawki}
To investigate the quality of the calibrations we employed $K_{\rm s}$-band
commissioning data from the VLT High Acuity Wide-field K-band Imager
\citep[HAWK-I,][]{p04,c06}. HAWK-I is a wide-field, near-IR camera
with four 2k x 2k Hawaii-2 detectors, covering a 7\farcm5 square field
at a pixel scale of 0\farcs106. Commissioning images of 30~Dor were
obtained on 2007 October 3, using $Y$, $J$, $K_{\rm s}$\/ and Br$\gamma$
filters. The HAWK-I exposures were dithered around the centre of 30~Dor to
build-up a complete mosaic for the region\footnote{See p.\,22 of the
December 2007 edition of The Messenger.}.
One $K_{\rm s}$-band HAWK-I array with a reasonable spatial overlap was selected for each
MAD image. These frames were then reduced and
analysed using the same methods as for the MAD images,
i.e. using PSF-fitting photometry. There were $\sim$50 stars from 2MASS with `AAA'
$J${$H$}$K_{\rm s}$\/ quality ratings in each of the HAWK-I frames. The
resulting $K_{\rm s}$\/ ZPs were applied to all of the objects found in the
image and a catalogue was created. A comparison of the HAWK-I
magnitudes with those from 2MASS for the calibration stars is shown in
Figure~\ref{phot_cf1}, with dispersions of $\pm$0.12, 0.16, and
0.21$^{\rm m}$ for Fields~1, 2 and 3, respectively.
These results provided a well-defined ZP between HAWK-I and 2MASS,
which was then used to bootstrap the MAD ZPs from stars overlapping in
the images. By comparing the instrumental magnitudes of 30 stars in
the MAD frames with their corresponding HAWK-I values, new ZPs were
calculated for each $K_{\rm s}$\/ MAD field. Note that the stars used were
hand-picked to be isolated and well within the dynamic range of the
MAD observations -- using fainter sources from the MAD data would have
pushed toward the sensitivity limits of the HAWK-I frames rather than
providing improved calibration of the MAD data. The internal
agreement of overlapping stars within the three HAWK-I frames used for
calibration of the three MAD fields was found to be better than
0.1$^{\rm m}$.
\begin{figure}
\begin{center}
\hspace{-0.85cm}\includegraphics[height=6cm]{fig4.pdf}
\caption{Residuals of the calibrated HAWK-I $K_{\rm s}$-band photometry compared with the
2MASS values. Field~1: open squares;
Field~2: open triangles; Field~3: crosses.}\label{phot_cf1}
\end{center}
\end{figure}
Stars that appear in the overlap regions of the MAD fields were used
as a double-check of the ZPs from the HAWK-I data. Fields~1 and 3
were in excellent agreement, while the ZP for Field~2 gave magnitudes
that were 0.07$^{\rm m}$ brighter, although still within the calculated
uncertainties; the ZP for Field~2 was corrected accordingly, resulting in
$K_{\rm s}$-band magnitudes on a common system.
Unfortunately there were no matching HAWK-I observations in the
{$H$}-band to employ a similar calibration method. Informed by the
behaviour of the $K_{\rm s}$-band results, we adopt the 2MASS ZPs from
Table~\ref{zero_points} for Fields~1 and 3, which have much smaller
standard deviations than found for Field~2. Instrumental magnitudes
were then compared for stars overlapping between Fields~1 and 3 as
done for $K_{\rm s}$; the resulting offsets were in excellent agreement with
those expected from the ZPs from Table~\ref{zero_points} (i.e.,
$\Delta${$H$}$\sim$0.16$^{\rm m}$). We then compared instrumental
magnitudes for stars overlapping between Fields~1 and 2. These
offsets were then used to calculate a new mean ZP for Field~2, which
is within the quoted uncertainties from the 2MASS comparison, but
has a much reduced error.
\subsubsection {PSF Fitting on Individual Frames}\label{indiv}
If the PSF of a star varies subtly between different frames (as is
likely with AO-corrected imaging) the resulting combined PSF may
broaden slightly and be noisier. To investigate if this effect was
the dominant source of our ZP uncertainties, we employed a second
photometric method to see if we could improve the calibrations. We
used PSF-fitting subtractions on the individual frames, then used
Dr.~Peter Stetson's {\sc daomatch} and {\sc daomaster} packages to
create a mean catalogue. This method potentially provides more
precise PSF fitting, leading to improved fidelity of the final
photometric catalogue. It also allows the user to set the number of
frames in which a source must be detected for its inclusion in the
final catalogue. This has the big advantage that objects in the
dithered regions (previously discarded when the combined image is
subset) can now be included, increasing the spatial extent of the
final source catalogue, and also leads to fewer rejections of stars
affected by cosmics or bad pixels in only one frame.
\begin{table}
\begin{center}
\caption{Photometric zero-points obtained from the individual
frame method. As in Table~\ref{zero_points}, the $K_{\rm s}$-band results are derived
from HAWK-I, with the {$H$}-band results from 2MASS. In both cases Field~2 was
calibrated using stars overlapping with Field~1.}\label{final_zero_points}
\begin{tabular}{lcccc}
\hline
Band & Field 1 & Field 2 & Field 3 \\
\hline
$K_{\rm s}$ & 23.40 $\pm$ 0.07 & 23.62 $\pm$ 0.08 & 23.59 $\pm$ 0.11 \\
{$H$} & 24.45 $\pm$ 0.14 & 24.37 $\pm$ 0.14 & 24.33 $\pm$ 0.08 \\
\hline
\end{tabular}
\end{center}
\end{table}
The uncertainties on the ZPs obtained from analysis of the individual
frames were (effectively) unchanged compared to those found from our previous best
efforts (see Table~\ref{final_zero_points}
cf. Table~\ref{zero_points}). This suggests that the uncertainties in
the magnitudes of the calibration stars are dominanting those from the
photmetric methods employed. A more tangible gain is the increased
number of sources detected due to the retention of the dithered
regions not observed in all frames (albeit with lower sensitivity).
Following application of the ZP calibrations, a comparison between the
combined-frame and individual catalogues was undertaken for each pointing/band,
finding mean offsets less than 0.05$^{\rm m}$ in all six instances.
The ZPs adopted for calibration of the MAD fields for the
final catalogue, created from the individual frames, are listed in
Table~\ref{final_zero_points}. In summary, the $K_{\rm s}$-band ZPs were calibrated using
HAWK-I, with Field~2 bootstrapped using overlapping stars in
Field~1. The {$H$}-band ZPs were calibrated using 2MASS, also with
Field~2 adjusted from comparisons with Field~1.
Figure~\ref{phot_cf2} compares the final MAD magnitudes with
those of the HAWK-I calibration stars, with the dispersions around the
mean quoted in Table~\ref{final_zero_points}.
\begin{figure}
\begin{center}
\hspace{-0.85cm}\includegraphics[height=6cm]{fig5.pdf}
\caption{Residuals of the calibrated MAD $K_{\rm s}$-band photometry from the individual frame
method compared with the HAWK-I values for hand-picked reference
stars (see text for details)}.\label{phot_cf2}
\end{center}
\end{figure}
As an external check on the MAD magnitudes we compared our
results for the 12 (visually-matched) overlapping stars in the
NICMOS observations from \citet{br01}. Nine of these stars were in Field~2 (with two
only observed in the {$H$}-band due to the dither pattern), with the
remaning three in Field~3. Accounting for the slight differences in
the photometric filters (the {\em HST} data were transformed to the
CTIO photometric system), the mean differences (MAD\,$-$\,{\em HST})
and standard errors are in excellent agreement
$\Delta${$H$}\,$=$\,$-$0.05\,$\pm$\,0.03 and
$\Delta$$K_{\rm s}$\,$=$\,$-$0.04\,$\pm$\,0.05 (std. err.).
\subsection {Colour-magnitude diagrams}
Colour-magnitude diagrams (CMDs) were created from catalogues from
both photometric methods (i.e. from the combined and the individual
frames), excluding objects within a radius of 2\farcs8 of the centre of R136,
where crowding/blending become dominant.
The individual frame method yields an extra 600-1,000 objects in the
CMDs compared to the analysis of the combined frames, mainly due to
the increase in the effective field-of-view. When CMDs from the
individual method are compared over the same spatial region as the
combined CMD, Fields~2 and 3 contain slightly fewer objects,
predominantly at fainter magnitudes, suggesting the individual
frame method results in slightly diminished sensitivity. In contrast,
Field~1 still contains fewer objects in the combined CMD, which is a
legacy of a light leak down one side of the images that affected some
of the MAD SD observations. Field~1 was the
most affected by this light leak in our observations, with the image quality parameters in
{\sc daophot} rejecting more sources in analysis of the combined frame
in this region than in the individual frames.
In addition to the inclusion of the dithered regions, the individual
frame method reduces the random star-to-star uncertainties in the
final catalogues. CMDs from the individual frame method are shown for
all three fields in Figure~\ref{CMDs}. The location of the
main-sequence is in good agreement for all three fields.
There is significant and variable extinction toward R136, with
\cite{a09} adopting a median extinction
of $A_V$\,$=$\,1.85 for massive stars (M\,$=$\,7-20\,M$_\odot$),
corresponding to $A_K$\,$\sim$\,0.21
\citep[assuming, for the purposes of an order of magnitude estimate of the
IR extinction, the standard galactic extinction law from][]{rl85} and
$A_H$\,$\sim$\,0.32 \cite[using the scaling from][]{i05}, yielding a
relatively small near-IR extinction term of E({$H$}$-$$K_{\rm s}$)\,$=$\,0.11.
Thus, the {$H$}$-$$K_{\rm s}$\/ colours from MAD will, in general, not be
differentially reddened by much more than the photometric errors.
CMDs were created for different radial bins using the combined
catalogue from all three fields and these show no significant
offset in the locus of the main sequence.
For Field~1 in Figure~\ref{CMDs} we include the unreddened
log$t$\,$=$\,3.0 isochrone (effectively illustrative of the zero-age
main sequence) from \citet{ls01}, adopting the tracks with the
metallicity relevant to that of the LMC \citep[i.e. those
from][]{s93}. Note that the offset from the MAD data is in agreement
with the reddening estimated above.
\begin{figure}
\begin{center}
\hspace{-0.8cm}\includegraphics[width=8.5cm]{fig6a.pdf}\\
\vspace{-0.35cm}\hspace{-0.8cm}\includegraphics[width=8.5cm]{fig6b_degraded.pdf}\\
\vspace{-0.35cm}\hspace{-0.8cm}\includegraphics[width=8.5cm]{fig6c.pdf}
\caption{CMDs from the mean catalogues created from the individual frames method,
excluding the central 2\farcs8 of R136. The dashed horizontal
lines mark the 50\% completeness level (cf. Figure~\ref{complete})
in the common regions included in all positional dithers. Overplotted in the
Field~1 panel is the {\em unreddened} youngest isochrone
(log$t$\,$=$\,3.0) from \citet{ls01}, with the marked points indicating,
from bright to faint magnitudes, initial masses of 60, 50, 25, 15, 9, 5, 3, and 2\,M$_\odot$.}\label{CMDs}
\end{center}
\end{figure}
\subsection{Photometric Completeness}
Completeness tests were undertaken in both bands for all three
fields. Using the {\sc starlist} and {\sc addstar} routines in {\sc
iraf}, 100 artificial stars with magnitudes varying from 14 to 24 were
distributed uniformly across each combined (and subset) image. As
before, PSF-fitting was performed, using the same settings and a {\sc
penny}{\small 2} model PSF which is allowed to vary across the field.
Using the known positions of the added stars, the number
recovered from the images within a two-pixel search radius was found.
If more than one object was found within that small radius, the object
with the smallest magnitude difference was considered the artificial
star. Stars with magnitude differences of greater than $\pm$0.5$^{\rm
m}$ were cut from the detected list, with the intention of excluding
mistaken matches, or artificial stars which have been superimposed on
genuine stars within the image (thus increasing its magnitude).
These tests were repeated 1,000 times, until 100,000 artificial stars
had been added to each image. This gives a ratio of $\sim$30 to 50
artificial stars for each observed object. The ratio of artificial
stars introduced to those detected was measured for each image, as
shown in Figure~\ref{complete}. Field 2, which does not
include R136 in the combined image, has a slightly more uniform
completeness at brighter magnitudes while going slightly deeper in both bands.
The 50\% completeness level in the $K_{\rm s}$-band frames is $K_{\rm s}$\,$=$\,19.45 in Field~1,
19.85 in Field~2, and 18.90 in Field~3, corresponding to initial main-sequence
masses of approximately 3.5, 2.8, and 4.7~M$_\odot$, respectively, when
compared to the youngest isochrone from \citet{ls01}.
For Fields~1 and 3 the completeness was also found for both bands as a
function of radius from the cluster (adopting 100 pixel radial bins),
an example of which is shown in Figure~\ref{rad_comp_Kp1}. These
radial completeness profiles are taken into account in construction
of the radial luminosity profile in Section~\ref{profiles}.
\begin{figure}
\begin{center}
\hspace{-0.85cm}\includegraphics[height=6cm]{fig7a.pdf}
\hspace{-0.85cm}\includegraphics[height=6cm]{fig7b.pdf}
\caption{{$H$}\/ ({\it upper panel}) and $K_{\rm s}$ ({\it lower}) completeness profiles for Fields~1, 2 and 3, as
indicated.}\label{complete}
\end{center}
\end{figure}
\begin{figure*}
\begin{center}
\includegraphics[scale=0.55]{fig8.pdf}
\caption{$K_{\rm s}$-band radial completeness for Field 1. Each annulus has a width of 100 pixels (2\farcs8).
`Radius~1' refers to the radius closest to the cluster core.}\label{rad_comp_Kp1}
\end{center}
\end{figure*}
\section{{\it Spitzer} YSO candidates}\label{ysos}
From observations with the {\em Spitzer Space Telescope}, GC09
presented a catalogue of potential young stellar objects (YSOs) in the
LMC. Prior to investigation of the luminosity profile of R136, we
first investigated the seven candidate YSOs from their catalogue that
are within the MAD fields. The {\em Spitzer} data were primarily
obtained under the auspices of the Surveying the Agents of a Galaxy's
Evolution (SAGE) LMC survey \citep{mm06}, but were also supplemented
by smaller, more targeted programmes from the {\it Spitzer} archive.
Following initial colour cuts to exclude the majority of stars and
background galaxies, GC09 investigated the nature of the candidate YSOs
via inspection of their morphologies and spectral energy
distributions, including comparisons at other wavelengths. This
resulted in five categories for the colour-selected {\it Spitzer}
sources: (1) evolved stars, (2) planetary nebulae, (3) background
galaxies, (4) diffuse sources, and (5) `definite', `probable' and
`possible' YSO candidates. This is a somewhat different approach to
the selection of candidate YSOs from the SAGE LMC survey by
\citet[][see further discussion by GC09]{w08}. \citet{v09} have since
compared the {\em Spitzer} data with archival {\em HST} H$\alpha$
images and new (seeing-limited) near-IR imaging, enabling more
detailed description of the local environments around 82 of the YSOs
from GC09.
The positional uncertainty on the {\it Spitzer} astrometry is
approximately $\pm$0\farcs2 (Dr~R.~Gruendl, priv. comm.), but in the
crowded environment of 30~Dor, one might expect that blended sources and
nearby gas might lead to greater uncertainties. Indeed, the typical
angular resolution of the {\it Spitzer} observations ranged
from 1\farcs7 to 2\farcs0 over the four InfraRed Array Camera (IRAC) bands
for the SAGE observations \citep{mm06}. Thus, it is not surprising that,
in the case of two of the `diffuse' sources from GC09, we find apparent counterparts
in the MAD images offset by a radius of approximately 1$''$.
In light of these positional uncertainties, there remains one
candidate from GC09 which does not have an obvious close counterpart
within a 2\farcs0 radius: the `diffuse' candidate {\it 053842.69
$-$690623.7}. There is a faint source ($K_{\rm s}$\,$=$\,18.1, {$H$}$-$$K_{\rm s}$\,$=$\,0.39)
at a distance of $\sim$0\farcs75, which is otherwise unremarkable; a brighter
counterpart can be ruled out. We note that of the four {\em
Spitzer}-IRAC bands, GC09 only reported magnitudes for this source at
4.5 and 8.0\,$\mu$m, suggesting a less robust detection than for the
other six sources discussed here which were detected in all four IRAC
bands.
\begin{figure}
\begin{center}
\includegraphics[width=8cm]{fig9.pdf}
\caption{Location of the six YSOs from \citet{g09} with counterparts
in the MAD images, overlaid on the $V$-band WFI image.}\label{yso_wfi}
\end{center}
\end{figure}
The remaining six YSO candidates are now discussed in turn, with their
locations overlaid on the $V$-band WFI image in Figure~\ref{yso_wfi}
to illustrate their positions relative to R136. All except {\it
053836.48 $-$690524.1} feature in the comparison by \citet{v09}, with
four also within the (seeing-limited) near-IR imaging from
\citet[][hereafter RBW98]{rbw98}. {\em Spitzer} spectroscopy of three
of these sources ({\it 053839.24 $-$690552.3}, {\it 053839.69
$-$690538.1} and {\it 053845.15 $-$690507.9}) was presented by
\citet{s09}, with each classified in their `PE Group', which are seen to display
polycyclic aromatic hydrocarbon (PAH) emission features combined with
fine-structure lines from ionised gas.
\subsection{\it{053836.48 $-$690524.1:}}
Reported as a `diffuse' source by GC09, this object is on the western
edge of the Field~2 images. Subset (5$''$\,$\times$\,5$''$) {$H$}\/ and
$K_{\rm s}$-band images are shown in Figure~\ref{yso1}, in which the cross in
the {$H$}-band image indicates the position from GC09 and the intensity
scaling is the same in both bands. The circle (centred on the {\it Spitzer}
position in the {$H$}-band image) indicates the typical angular resolution
of IRAC in the SAGE survey, i.e. 1\farcs7 at 3.6\,$\mu$m \citep{mm06}.
Just below the {\it Spitzer} position in the {$H$}-band
image is a small cluster of bad pixels -- the likely YSO counterpart is
to the east and is more clearly seen in the $K_{\rm s}$-band image.
Apart from the obvious asymmetry in the $K_{\rm s}$-band image, we were
somewhat cautious to match the MAD object to the GC09 source.
However, \citet{kk07} also report a candidate YSO (their
`30\,Dor-15', also from IRAC observations) that is only
$\sim$0\farcs33 from the MAD counterpart, which has $K_{\rm s}$\,$=$\,16.24,
{$H$}\,$-$\,$K_{\rm s}$\,$=$\,2.08.\footnote{We note the candidate YSO `30\,Dor-22' from
\citet{kk07} is, on the basis of their published astrometry, Mk\,34 \citep{m85}, a
well-studied WN star. We are puzzled that Kim et al. mention this as
one of the sources studied by \citet{br01}, as it does not feature in their NICMOS fields.}
\begin{figure}
\begin{center}
\includegraphics[height=3.75cm]{fig10a.pdf}\hspace{0.1cm}\includegraphics[height=3.75cm]{fig10b.pdf}
\caption{053836.48 $-$690524.1: {$H$}- and $K_{\rm s}$-band 5$'' \times$5$''$ images
(left- and right-hand panels, respectively), with north at the top,
east at the left. The black cross marks the location of the {\it
Spitzer} position for the YSO candidate, with the black circle representing
the angular resolution of the IRAC data (1\farcs7 at 3.6\,$\mu$m).}\label{yso1}
\end{center}
\end{figure}
\subsection{\it{053838.35 $-$690630.4:}}
Classified as a `diffuse source' by GC09, there is a potential counterpart
($K_{\rm s}$\,$=$\,14.37 and {$H$}$-$$K_{\rm s}$\,$=$\,0.35) in the MAD images at a distance
of 1$''$ (Figure~\ref{yso2}). The source appears to have an associated
`plume' of nebulosity which spirals to the south-west for at least
0\farcs5. This nebulosity can also be seen in the NICMOS image from
\citet[][their 30Dor-NIC01 frame]{br01}, but did not feature in their discussion.
The 2MASS $J$-band magnitude quoted by GC09
for this source is 13.93 (2MASS source: 053838.5-6906297).
Overplotting the 2MASS catalogue on the MAD image reveals reasonable
astrometric agreement (0\farcs2) considering the limited angular
resolution of 2MASS, with upper limits of {$H$}\,$>$\,12.88 and $K_{\rm s}$\,$>$\,12.05 (i.e. `U' rated).
This source also corresponds to IRSW-98 from RBW98
\citep[which, in turn, is component `a' of the knot comprising source W4
from][]{rrg92}, who found $K_{\rm s}$\,$=$\,14.02 and {$H$}$-$$K_{\rm s}$\,$=$\,0.45.
\citet{v09} note that this YSO is in a marginally-resolved H~\2 region,
suggesting that the line contribution (at least in the $K_{\rm s}$-band) is likely
contributing to the brighter magnitude from RBW98 compared to the MAD results.
\subsection{\it{053839.24 $-$690552.3:}}
The most likely YSO counterpart is at $\alpha$\,$=$\,05$^{\rm
h}$\,38$^{\rm m}$\,39\fs28, $\delta$\,$=$\,$-$69$^\circ$\,05$'$\,52\farcs60 (J2000,
$\sim$0\farcs45 from the {\it Spitzer} position) and
is the most immediately impressive counterpart in the MAD images
(Figure~\ref{yso3})\footnote{The source is on the edge of the
region common to all {$H$}- and $K_{\rm s}$-band images, hence the apparent `banding'
where the sensitivity differs due to the science-frame dithers.}. We find
$K_{\rm s}$\,$=$\,14.71 and {$H$}$-$$K_{\rm s}$\,$=$\,0.42 for the star, as compared to
$K_{\rm s}$\,$=$\,13.91, {$H$}$-$$K_{\rm s}$\,$=$\,0.61 from RBW98 (their source IRSW-118).
The images from RBW98 were taken in typical seeing of 1$''$ so
the two sources visible a the centre of Figure~\ref{yso3} will be
strongly blended, leading to the brighter magnitude.
Particularly striking is the apparent bow-shock, most easily seen in the
$K_{\rm s}$-band image. This is almost, but not quite, aligned with the core of
R136 approximately 19\farcs5 away. Also of note in the same direction
is R134/Brey~75 \citep{b81}, classified as WN6(h) by \citet{cs97}, at
a distance of only 8$''$. To date, nothing is known regarding the
spectral type of the two bright stars at the centre of the image
(separated by $\sim$0\farcs33). This region clearly warrants more
detailed investigation in the context of small-scale triggered star
formation.
\subsection{\it{053839.69 $-$690538.1:}}
Classified as a `definite YSO' by GC09, this object appears
as a somewhat extended or embedded source in the {$H$}-band MAD image
(Figure~\ref{yso4}), with $K_{\rm s}$\,=\,13.75 and {$H$}$-$$K_{\rm s}$\,=\,3.37.
This source corresponds to IRSW-127 from RBW98 \citep[W9
from][]{rrg92}, for which they found $K_{\rm s}$\,$=$\,13.91 and
{$H$}$-$$K_{\rm s}$\,$=$\,2.85.
The bright star 1\farcs2 to the south is P733/S206, for which
$K_{\rm s}$\,=\,14.40 and {$H$}$-$$K_{\rm s}$\,$=$\,0.16, suggesting it as a massive star
\citep[cf. the intrinsic colours from][]{mp06}.
The strong contrast in the colours of the two
stars is immediately obvious from the composite-colour image from
RBW98 (their Figure~1), although they appear slightly blended.
Blending could perhaps account for the photometric differences (which are in
the expected direction), but RBW98 employed PSF-fitting methods so
should be relatively robust to such effects at this separation --
perhaps our efforts are somewhat limited by the apparent asymmetric
flux to the south of the star.
\subsection{\it{053843.52 $-$690629.0:}}
Classified as a `possible YSO' by GC09, the nearest spatial match is a source with
$K_{\rm s}$\,=\,17.15 and {$H$}$-$$K_{\rm s}$\,$=$\,1.01, as indicated in Figure~\ref{yso5}. The adjacent
object (0\farcs25 to the northeast) is P1064/S696, with $K_{\rm s}$\,$=$\,16.87 and
{$H$}$-$$K_{\rm s}$\,$=$\,0.18, and for which the spectral type is unknown. This YSO source is not
within the region observed by RBW98 but, interestingly, is noted by \citet{v09} as being
in a dark cloud and as comprising multiple YSOs in the {\em Spitzer} PSF,
one with an optical counterpart in the H$\alpha$ {\em HST} images, and
one without.
\subsection{\it{053845.15 $-$690507.9:}}
Classified as a `definite YSO' by GC09, their published position is
approximately 0\farcs4 north of P1222/S116/IRSN-101 \citep[][RBW98]{p93,s98}, which was
classified as O3-6~V by \citet{wb97} from ground-based spectroscopy, later
revised to O9~V(n)p by \citet{wal02} from {\em HST} observations;
5$''\times$5$''$ images of the region are shown in Figure~\ref{yso6}.
Note that P1222 was not used in the astrometric calibration of the
MAD frames and yet its recovered position agrees with Skiff's astrometry
to better than two pixels.
This source lies within the dense nebular region referred to as
`Knot~1' by \citet{wb86}, highlighted as young massive
stars just emerging from their natal cocoons and imaged at optical and
near-IR wavelengths with {\em HST} by \citet{wal99,wal02}.
The relatively bright
{\em Spitzer} magnitudes quoted by GC09 are in keeping with P1222 as
the most plausible counterpart. Indeed, the spectral energy
distribution shown in Figure~13 of GC09 reveals increased magnitudes
bluewards of 1\,$\mu$m, as well as redwards, consistent with a massive
star with an IR excess. From the MAD images we find $K_{\rm s}$\,$=$\,14.31
and {$H$}$-$$K_{\rm s}$\,$=$\,0.13, as compared to $K_{\rm s}$\,$=$\,14.45 and
{$H$}$-$$K_{\rm s}$\,$=$\,0.03 from the {\em HST}-NICMOS observations of Knot~1
by \citet{br01}.
\citet{v09} note the source as residing in a resolved H~\2 region and, on the
basis of the H$\alpha$ luminosity (uncorrected for extinction) provide a lower
bound on the inferred spectral type of the YSO of B0.5; evidently from the
classification of \citeauthor{wb97}, the star is slightly hotter.
\begin{figure}
\begin{center}
\includegraphics[height=3.75cm]{fig11a.pdf}\hspace{0.1cm}\includegraphics[height=3.75cm]{fig11b.pdf}
\caption{053838.35 $-$690630.4: 5$''$\,$\times$\,5$''$ {$H$}- and $K_{\rm s}$-band images.}\label{yso2}
\includegraphics[height=3.75cm]{fig12a.pdf}\hspace{0.1cm}\includegraphics[height=3.75cm]{fig12b.pdf}
\caption{053839.24 $-$690552.3: {$H$}- and $K_{\rm s}$-band. Note the
particularly striking bow-shock feature in the $K_{\rm s}$\/ image.}\label{yso3}
\includegraphics[height=3.75cm]{fig13a.pdf}\hspace{0.1cm}\includegraphics[height=3.75cm]{fig13b.pdf}
\caption{053839.69 $-$690538.1: {$H$}- and $K_{\rm s}$-band.}\label{yso4}
\includegraphics[height=3.75cm]{fig14a.pdf}\hspace{0.1cm}\includegraphics[height=3.75cm]{fig14b.pdf}
\caption{053843.52 $-$690629.0: {$H$}- and $K_{\rm s}$-band.}\label{yso5}
\includegraphics[height=3.75cm]{fig15a.pdf}\hspace{0.1cm}\includegraphics[height=3.75cm]{fig15b.pdf}
\caption{053845.15 $-$690507.9: {$H$}- and $K_{\rm s}$-band.}\label{yso6}
\end{center}
\end{figure}
\subsection{Spatial distribution and masses of YSOs}
It is not surprising that five of the six YSO candidates discussed
here lie to the north and west of R136 (Figure~\ref{yso_wfi}). This is
the location of the `second generation' of massive stars
\citep{wal99,wal02}, with significant molecular material still
remaining \citep[e.g.][]{w78,j98}. The sixth source ({\it 053843.52
$-$690629.0}) is spatially distinct from the regions of triggered star
formation and is the most ambiguous of the YSOs discussed here;
with M$_K$\,$\sim$\,$-$1.4, it is also likely to be the lowest mass object.
The absolute $K_{\rm s}$-band magnitudes of our four brightest YSO
counterparts are all bright, with M$_K$\,$\sim$\,$-$3.75-4.75,
suggesting that they are {\it bona fide} candidates for massive O- or
early B-type stars, depending on the degree of obscuration/reddening.
The most intriguing in this context is the very red source {\it
053839.69 $-$6905381} which, given the uncertainties of the
{\em Spitzer} astrometry, would certainly benefit from ground-based
spectroscopy to elucidate its nature and its relationship with P733 nearby.
\citet{r09} have recently reported a dense molecular cloud toward this
source (their IRSW-127) suggesting that it may well be a massive star
still heavily enshrouded in its natal cocoon.
\section{Radial Luminosity Profile of R136}\label{profiles}
We now compare the radial luminosity profile for R136 from the MAD
data, constructed using a combination of integrated-light measurements and
star counts, with the optical {\em HST} results.
\subsection{Integrated Light Profiles}
Integrated-light measurements were used to investigate the innermost
part of the radial luminosity profile in Fields~1 and 3. These
sum the flux within annuli around the core, where the
radius of each annulus is defined such that the total flux in each is
approximately constant.
An {\sc idl} program was written specifically for this purpose, with standard
{\sc daophot} routines in {\sc iraf} also used as a comparison. The
{\sc idl} program calculates the integrated light by re-sampling each
image as a function of azimuth ($\phi$) and radius from the cluster
centre, to enable a smoother, more robust calculation at smaller radii
than possible with the (albeit small) sampling of the pixels -- in
practice the results are essentially the same as those obtained with
{\sc iraf}. These measurements adopt a constant sky value, which is
defined well away from the cluster core.
To estimate the uncertainties in the integrated-light profile, each
annulus is split into azimuthal sections and the density is calculated
in each; the standard deviation of the results around each annulus
then provide an estimate of the statistical uncertainties, provided the
cluster is symmetric. However, visual inspection of the images
shows that the cluster is not symmetric, particularly at larger radii
(see Section~\ref{azvar}), suggesting that the variations in the uncertainties
are shaped by both statistical errors and asymmetric variations of the cluster.
The surface brightness profiles were then fit with \citet[][EFF]{eff} cluster
profiles of the form:
\begin{equation}
\mu(r)=\mu_0 (1+\frac{r^2}{a^2})^{-\gamma/2},
\end{equation}
where $a$ is the scale radius that defines where the flat inner part
(the core) turns into a power-law profile with an index of
$-\gamma$. The results are given in Table~\ref{eff_tab}, both in terms of angular extent on
the sky and in parsecs (adopting a distance modulus to the LMC of
18.5). The profile fits are shown in Figure~\ref{eff_fits1}, in which the lower three profiles
are offset (in steps of two magnitudes) for clarity.
There is hardly any signature of a core, with the surface brightness
profile fit by a power-law function over the full range of
the data. The scale radius ($a$) obtained is effectively
$\sim$\,0.1$''$ (0.025\,pc) which is, in practice, the resolution limit
of these data; i.e. 0.1$''$ provides an upper limit to the core
radius\footnote{Note that the core radius, $r_{\rm core}$, is usually
defined as the radius where the surface brightness profile drops to
half its central value. For an \citeauthor{eff} profile it follows
that $r_{\rm core}=a\,(2^{2/\gamma}-1)$. For the values of $\gamma$ we
find here, $r_{\rm core}\approx1.5\,a$.}. Both the results for the core and
the slope of the power-law fit (i.e. $\gamma$\,$\sim$\,1.6) are in good
agreement with past optical {\em HST} results \citep{c92,h95} and the new
NICMOS results \citep{a09}.
\begin{table*}
\begin{center}
\caption{Structural parameters for R136 from EFF fits to the MAD integrated-light profiles. The italicised
results are from fits to the combined luminosity profiles (see Section~\ref{combined}). For
convenience, results are quoted in direct observables and their physical size in pc.}\label{eff_tab}
\begin{tabular}{lcccc}
\hline
Field & $\mu_0$ [mag arcsec$^{-2}$] & $a$ [$''$] & $\gamma$ & $r_{\rm c}$ [$''$] \\
\hline
H F1 & 6.37\,$\pm$\,0.12 & 0.13\,$\pm$\,0.03 & 1.67\,$\pm$\,0.15 & 0.15\,$\pm$\,0.04\\
& {\em 6.36}\,$\pm$\,{\em 0.12} & {\em 0.12}\,$\pm$\,{\em 0.02} & {\em 1.57}\,$\pm$\,{\em 0.05} & {\em 0.14}\,$\pm$\,{\em 0.03} \\
H F3 & 6.59\,$\pm$\,0.11 & 0.12\,$\pm$\,0.03 & 1.58\,$\pm$\,0.13 & 0.14\,$\pm$\,0.03\\
& {\em 6.59}\,$\pm$\,{\em 0.11} & {\em 0.12}\,$\pm$\,{\em 0.02} & {\em 1.57}\,$\pm$\,{\em 0.05} & {\em 0.13}\,$\pm$\,{\em 0.02}\\
K F1 & 6.35\,$\pm$\,0.14 & 0.13\,$\pm$\,0.04 & 1.58\,$\pm$\,0.15 & 0.15\,$\pm$\,0.04\\
& {\em 6.35}\,$\pm$\,{\em 0.14} & {\em 0.12}\,$\pm$\,{\em 0.02} & {\em 1.48}\,$\pm$\,{\em 0.06} & {\em 0.15}\,$\pm$\,{\em 0.03}\\
K F3 & 6.39\,$\pm$\,0.10 & 0.14\,$\pm$\,0.03 & 1.63\,$\pm$\,0.14 & 0.16\,$\pm$\,0.04\\
& {\em 6.38}\,$\pm$\,{\em 0.11} & {\em 0.12}\,$\pm$\,{\em 0.02} & {\em 1.56}\,$\pm$\,{\em 0.06} & {\em 0.15}\,$\pm$\,{\em 0.03}\\
\hline
&&&&\\
\hline
Field & $\mu_0$ [mag pc$^{-2}$] & $a$ [pc] & $\gamma$ & $r_{\rm c}$ [pc] \\
\hline
H F1 & 3.293\,$\pm$\,0.121 & 0.031\,$\pm$\,0.008 & 1.67\,$\pm$\,0.15 & 0.035\,$\pm$\,0.009\\
& {\em 3.282}\,$\pm$\,{\em 0.121} & {\em 0.028}\,$\pm$\,{\em 0.005} & {\em 1.58}\,$\pm$\,{\em 0.05} & {\em 0.033}\,$\pm$\,{\em 0.006} \\
H F3 & 3.517\,$\pm$\,0.112 & 0.028\,$\pm$\,0.007 & 1.58\,$\pm$\,0.13 & 0.034\,$\pm$\,0.008\\
& {\em 3.509}\,$\pm$\,{\em 0.112} & {\em 0.027}\,$\pm$\,{\em 0.005} & {\em 1.57}\,$\pm$\,{\em 0.06} & {\em 0.032}\,$\pm$\,{\em 0.006} \\
K F1 & 3.278\,$\pm$\,0.136 & 0.031\,$\pm$\,0.009 & 1.58\,$\pm$\,0.15 & 0.037\,$\pm$\,0.010\\
& {\em 3.269}\,$\pm$\,{\em 0.136} & {\em 0.029}\,$\pm$\,{\em 0.006} & {\em 1.48}\,$\pm$\,{\em 0.06} & {\em 0.035}\,$\pm$\,{\em 0.007} \\
K F3 & 3.316\,$\pm$\,0.105 & 0.033\,$\pm$\,0.008 & 1.63\,$\pm$\,0.14 & 0.039\,$\pm$\,0.010\\
& {\em 3.305}\,$\pm$\,{\em 0.105} & {\em 0.030}\,$\pm$\,{\em 0.006} & {\em 1.56}\,$\pm$\,{\em 0.06} & {\em 0.036}\,$\pm$\,{\em 0.007} \\
\hline
\end{tabular}
\end{center}
\end{table*}
\begin{figure}
\begin{center}
\hspace{-0.9cm}\includegraphics[width=8.75cm]{fig16.pdf}
\caption{EFF fits to the MAD integrated-light profiles, with the $K_{\rm s}$- and {$H$}-band (Field~3) profiles
shifted to fainter magnitudes for clarity.}\label{eff_fits1}
\end{center}
\end{figure}
\subsection{Star Counts}\label{direct}
Star counts in annuli around the core were used to calculate the
radial luminosity profile in the regions beyond $r$\,$=$\,2\farcs8.
Stars were binned such that a similar number of objects were included
per annulus, as suggested by \citet{au05}. The stars were binned as a
function of radius and magnitude and completeness terms were
calculated. As the core is off-centre in each frame, area correction
terms were applied to correct the densities if part of a given annulus
extended past the edge of an image.
The radial completeness tests (e.g. Figure~\ref{rad_comp_Kp1}) reveal,
not unsurprisingly, that the first couple of annuli beyond
$r$\,$=$\,2\farcs8 are also strongly affected by crowding, and are
thus subject to greater uncertainties than in the outer regions. The
50$\%$ completeness level in the third radial bin was adopted as the
faint limit for construction of the profiles from the star counts,
corresponding to a faint magnitude limit in each field in the range of
18.5-18.8$^{\rm m}$, equating to a main-sequence mass of
$\sim$5\,M$_\odot$. The innermost bins were not included in
construction of the final combined profile.
\subsubsection{Azimuthal Density Variations}\label{azvar}
To investigate the apparent asymmetry in R136 more quantitatively, we
calculated the azimuthal density variation beyond the central 2\farcs8.
An {\sc idl} program was used to split the images into angular slices,
each containing approximately equal numbers of stars (after
completeness corrections were applied). Similar techniques have been
used recently to investigate the structure and asymmetries seen in
lower-mass Galactic clusters \citep[e.g.][]{gm05,gm09}.
Due north from the cluster core was set as $\phi$=0$^\circ$,
increasing anti-clockwise (increasing towards the east), with the
outer radius set by the distance from the cluster core to the northern
edge of Field~1. This enables a near complete azimuthal sweep in the
combined {$H$}-band images of Fields~1 and 3 (i.e. not the full extent of
those shown in Figure~\ref{obs}). An iterative routine was used to
define an azimuthal slice in each image (increasing by 1$^\circ$ each
loop), which summed the star counts in that slice and corrected for
the incompleteness (down to 50\%). Once 200 stars are obtained, the
routine determines the area covered by each slice, calculates the
density, and then moves on to define the next slice.
The azimuthal variation around R136 (with $r_{\rm max}$\,=\,28\farcs0,
$\equiv$\,1,000 pixels) is clearly seen in
Figure~\ref{angular_density}, demonstrating that this region is far
from symmetric, with a local minimum around $\phi$\,$=$\,90$^\circ$.
Note that the {$H$}-band densities are larger as a consequence of the
deeper images compared to the $K_{\rm s}$-band observations (see
Figure~\ref{complete}). The exact densities are sensitive to the
completeness threshold employed, so similar calculations were
undertaken for completeness cuts of 30, 40, and 60\% (with the results
for 40\% shown in Figure~\ref{az2}). While there are variations
in the exact densities as a function of $\phi$ (partly because the
different depths yield slices which vary in size in terms of $\phi$),
the overall trend remains the same.
Adopting a distance modulus of 18.5 to the LMC, $r$\,=\,28\farcs0
corresponds to a projected radius of 6.8\,pc. Similar trends in the
azimuthal profiles are also seen adopting outer radii of 600 and 800
pixels (4.1 and 5.4\,pc, respectively). At
smaller radii incompleteness effects become more significant
(cf. Figure~\ref{rad_comp_Kp1}) and it is harder to characterise the
asymmetry meaningfully. Superficial inspection of Figure~2 from
\citet{mh98} suggests similar asymmetries in the inner region, with
luminous stars appearing to be more numerous in the eastern half of
the central 4\farcs55\,$\times$\,4\farcs55 {\em HST} image. However,
we also note that the most luminous, massive stars will likely have a
different relaxation time to the lower-mass population.
\begin{figure}
\begin{center}
\hspace{-0.85cm}\includegraphics[height=6cm]{fig17.pdf}
\caption{Azimuthal density profiles for a maximum radius of 28\farcs0 from the core of R136 and
a completeness cut-off of 50\%, demonstrating that the cluster is not relaxed over a $\sim$7\,pc scale.
Due north from the cluster centre corresponds to $\phi$ = 0${^\circ}$.}\label{angular_density}
\hspace{-0.85cm}\includegraphics[height=6cm]{fig18.pdf}
\caption{As Figure~\ref{angular_density}, but with a completeness cut-off of 40\%.}\label{az2}
\end{center}
\end{figure}
\subsection{Combined Radial Profile}\label{combined}
The overlap region between the profiles from the two methods was used
to convert (and normalise) the star counts into surface brightness
-- a reasonable approach provided there is no evidence of mass
segregation in the cluster \citep[as claimed by][although Brandl et al. 1996
argued to the contrary]{h95}. The average
offset in the overlap region was found from interpolation between the
two profiles, re-sampled to an evenly sampled grid of radial bins to
avoid a bias toward smaller radii (at which the bins are smaller) and
excluding the innermost bin from the star counts profile (to be less
sensitive to completeness corrections). The offset was then used to
convert the star counts into a magnitude density, yielding a combined
profile.
EFF fits to the combined profile are shown in
Figures~\ref{eff_fits_f1} and \ref{eff_fits_f3}, with the fit
parameters given in italics in Table~\ref{eff_tab}. Meaningful errors
on the star counts are hard to estimate due to the difficulty
of attributing uncertatinties to the completeness corrections
(i.e. blending/crowding effects) and the area corrections
(i.e. cluster asymmetries), so we adopt a conservative error of
$\pm$10\% of the density values for each bin, which are then scaled as per the
profile into a magnitude density. Note that the slopes of the fits are
nearly identical to those from analysis of the integrated-light
profile, although the results are very slightly shallower.
\begin{figure}
\begin{center}
\hspace{-0.9cm}\includegraphics[width=7.8cm]{fig19.pdf}
\vspace{-0.425cm}\caption{EFF fits to the combined MAD profiles for Field~1 (with the $K_{\rm s}$-band profile offset to fainter
magnitudes for clarity).}\label{eff_fits_f1}
\vspace{-0.25cm}\hspace{-0.9cm}\includegraphics[width=7.8cm]{fig20.pdf}
\vspace{-0.425cm}\caption{EFF fits to the combined MAD profiles for Field~3 (with the $K_{\rm s}$-band profile offset to fainter
magnitudes).}\label{eff_fits_f3}
\vspace{-0.25cm}\hspace{-0.9cm}\includegraphics[width=7.8cm]{fig21.pdf}
\vspace{-0.425cm}\caption{Profile fits to the {\em HST} optical data from
\citet{mg03}, with the $I$-band profile offset to fainter magnitudes.}\label{eff_fits3}
\end{center}
\end{figure}
There are striking differences between the EFF fits to the MAD data
and the optical results from \citet{mg03}. We also fitted EFF
profiles to the data from \citeauthor{mg03}, with it truncated to the
outer radius of the MAD data. However, without measurements closer
to the core it is difficult to obtain a meaningful comparison for
$\mu_0$ and $a$. Their inner data points suggest a somewhat steeper
turnover, with \citeauthor{mg03} reporting
$\gamma$\,$=$\,2.43\,$\pm$\,0.09 for their inner EFF component, but
the integrated-light MAD measurements at $r$\,$<$\,1$''$ rule out such
a slope in the near-IR. Most relevantly, fits to the slope of the
optical data yield $\gamma$\,$=$\,1.80\,$\pm$\,0.10 in both the $V$
and $I$ bands, in reasonable agreement with the MAD data, as
illustrated in Figure~\ref{eff_fits3}\footnote{These are the original data from \citeauthor{mg03}.
\citet{mvm05} later identified a magnitude offset of $-$0.715 (i.e. a
brighter profile) compared to Mackey \& Gilmore, who had masked out
bright stars.}. Note the presence of the bump (or dent) noted by both \citet{m93} and
\citet{mg03}, at a radius of approximately 10$''$. For comparison, \citet{mvm05} found
$\gamma\,=\,$2.05 from fits to the full range of the {\em HST}
data\footnote{\citet{mvm05} find $\gamma_{\rm 3D}\,=\,$3.05, which is
the slope of the 3D density profile, where $\gamma_{\rm
3D}\,=\,\gamma\,+\,$1 because of projection).}
A value of $\gamma$\,$=$\,1.65\,$\pm$\,0.08 was also found for the
slope of the number density profile for stars brighter than
M$_V$\,$=$\,$-$5 by \citet{mds94}. The extension of the power-law out
to $\sim$\,10\,pc makes R136 a good example of clusters with an extended
halo -- not all young clusters have such a halo, as illustrated by
\citet{mds94} through a comparison with NGC~3603, which
lacks bright stars beyond $\sim$\,1 pc. \citet{jma01} shows several
extra-galactic examples of clusters with and without an extended
halo. The origin of such halos in some clusters compared to those without
remains unclear at present.
\subsection{Discussion}\label{discussion}
\subsubsection{Structural parameters of R136}
In summary, we find slopes of $\gamma\,\le\,$2 from EFF fits to the
near-IR luminosity profile of R136. If extrapolated to infinity such
a profile would have infinite mass, so this can not extend to very
large radii as the bulk of the mass would then be far from the core.
It remains to be seen what the origin of these very shallow profiles
is, although \citet{l04} found that shallower profiles were more
common in the younger members of their extragalactic cluster sample,
and that $\gamma$ appears to increase with age. The development of a
core and the reduction of the central density with age is likely the
result of mass loss by stellar evolution and heating by binaries in
the centre \citep{tp00}.
Due to the very young age of R136 its density profile could be related
to the formation process of the cluster, i.e. while it is still in its earliest
stages of evolution we might see the imprint of the parent molecular
cloud. Molecular clouds are approximately isothermal spheres in which,
with $\gamma$\,$=$\,1 (i.e. $\gamma_{\rm 3D}$\,$=$\,2), the density
scales as $\rho$\,$\propto$\,$r^{-2}$ \cite[e.g.][]{m84}.
For example, \citet{hs06} found that the Orion Nebula Cluster (with an age of $\sim$1\,Myr) can be
well approximated by a power-law profile with $\gamma\,=\,$1.
Alternatively, the dissipationless collapse of a (gas free) stellar
system also results in a power-law density profile, and possibly a
core. The size of the core and the slope of the density profile
depends mainly on the virial temperature in the initial configuration
\citep{m84}. For a cold collapse, i.e. zero initial velocities of the
stars, \citeauthor{m84} found $\sim\,$1.5$\,<\,\gamma\,<\,\sim\,$2, with a core size of almost
zero. The `warmer' the initial conditions, the steeper the density
profile and the larger the size of the core in the final
configuration. So the observed density profile in R136 could
also be the result of a cloud collapse, with star formation occurring
in the early phases of the collapse, such that the dynamical
interactions between the stars determine the properties of the
cluster.
\citet{h95} found no evidence for mass segregation in R136.
\citet{a09} also ruled it out for $r$\,$>$\,3\,pc ($\equiv$12\farcs4),
finding no evidence for the flattening of the IMF below 2\,M$_\odot$
reported by \citet{s00}. An early AO study of one 12\farcs8\,$\times$\,12\farcs8
quadrant of R136 by \citet{b96} sampled a similar mass range to the MAD
data, and found evidence of mass segregation above 12\,M$_\odot$.
Whether the MAD data shows effects of the more massive stars being
preferentially in the core will be addressed elsewhere, but
if evidence for mass segregation were found in the
inner 3\,pc, then the core radius derived here would be somewhat smaller
than the 'primordial' radius.
\subsubsection{Structure at larger radii}
Evidence for a second component in the near-IR profiles is less
compelling than in the optical data, and we do not invoke a
second component in our EFF fits. There is tentative evidence in
Field~1 for a `dent' at comparable radius to the
optical data (Figure~\ref{eff_fits_f1} cf. Figure~\ref{eff_fits3}), but the data for
Field~3 are fit more cleanly by a single-component profile
(Figure~\ref{eff_fits_f3}). Unfortunately the MAD data do not lend
themselves to analysis of the integrated light out to larger radii.
While the `stitching together' of the two parts of the profile can
influence the appearance of the combined profile, exclusion of one or
two bins in the overlap region when combining them has very
little influence on the EFF results. Indeed, now armed with knowledge of the
minimal core, the EFF fits to the optical data are well matched at larger
radii.
The lack of a second component is interesting since theoretical
explanations for bumps (or halos) around young dense clusters
\citep[e.g.][]{bg06} should be independent of the wavelength in which
the profile is constructed. The temptation is to attribute the past
(two-component) results to variable extinction in the optical images.
From the optical image of R136 (Figure~\ref{obs}) there is a notable
`void' to the north-east (running from north-west to south-east) at
$r$\,$\sim$\,10$''$, suggesting that differential extinction could be
a prime suspect. We also note that more prosaic reasons (using
circular annuli to determine the luminosity profiles in elliptical
clusters) are now also being suggested to account for apparent breaks
in density profiles in other clusters \citep{p09}.
In this context, the asymmetric structure seen in
Figure~\ref{angular_density} must contribute strongly -- almost
certainly leading to the differences seen in the outer regions in
Figures~\ref{eff_fits_f1} and \ref{eff_fits_f3}. Further support of
this hypothesis is provided by inspection of the original Wide Field
Planetary Camera~2 (WFPC2) images used by
\citeauthor{mg03}. R136 was located slightly off-centre in the Planetary
Camera detector, with the Wide-Field Cameras primarily sampling the
region east of R136, i.e. Field~1 observed with MAD.
Interestingly, \citet{a09} only fit their EFF profile to the inner 2\,pc
($\equiv$8\farcs25), noting in the caption to their Figure~12 that
the presence of individual bright stars introduces `the jitter in the
surface brightness profile', hinting at the arguments that we advance here.
Such deviations from symmetry are not surprising in such a massive, intricate,
multi-population and, as yet, seemingly unrelaxed star-formation
region. This point is reinforced by the identification of (at least)
five distinct populations in 30~Dor by \citet{wb97}, three of which
are sampled by the MAD observations: the central `Carina Phase' (rich
in early O-type stars and including R136), an older `Scorpius OB1 Phase'
of early-type supergiants throughout the central field, and a young (likely triggered)
`Orion Phase' to the north and west, mostly embedded in the gaseous
filaments visible in Figure~\ref{obs} and partially observed by Field~2.
The stars associated with these populations in the region immediately
around R136 are neatly illustrated by Figures~3 and 4 from \citet{w86}
-- their distribution is far from uniform, illustrative of the
asymmetries shown in Figure~\ref{angular_density}.
The twin populations around R136 invite the question of their
formation history. Is the Sco~OB1 Phase simply the remaining `field'
population, dispersed from a previous star-formation event? Are the
younger, more massive Carina Phase stars formed locally, or are they
ejected members of R136? The answer to the latter question is most
likely a combination of both. Without comprehensive stellar and gas
dynamics of the
different populations in 30~Dor we are limited to speculation at the
present time, but this is one of the principal motivations for the
VLT-FLAMES Tarantula Survey \citep{f2}, which has obtained multi-epoch
optical spectroscopy for over 1,000 stars and will address
these points.
The formation history of these `halo' stars goes to the heart of the
theory of infant mortality, where deviations from EFF profiles
in M82-F, NGC\,1569-A and NGC\,1705-1 were argued by \citet{bg06} to
arise from rapid gas removal in clusters undergoing violent relaxation.
Note that the {\em HST} profile from \citet{mg03} extends
beyond a radius of 60$''$, thus also sampling the triggered
generation as well as the halo around R136, i.e. some of the excess
light reported by \citeauthor{mg03} (and, by inference, by \citeauthor{bg06})
is likely attributable to ejected stars, but that triggered star formation
also contributes.
To echo the sentiment of \citet{wb97}, the central 2$'$ of 30~Dor
observed with MAD corresponds to a physical size of $\sim$30\,pc. If
one projects this to a distance of 10\,Mpc, the same structures would
all be contained within $\sim$0\farcs5. This reinforces the need to
take potential cluster asymmetries and triggered star-formation into
account when interpreting the radial luminosity profiles of distant
unresolved clusters.
\bigskip
The MAD data have provided us with a truly unique view of the central region of
30~Doradus. In the broader context, the delivery of impressive image quality and
stability across wide fields (compared to classical AO observations)
is very encouraging in advance of future multi-laser, MCAO systems on
8-m class telescopes, and the development of wide field,
AO-corrected, imagers and spectrometers for the E-ELT.\\
\noindent {\sc acknowledgments:}
The data presented here were obtained during the MAD Science
Demonstration campaign (package name MADD-SD-EV, ID 96408).
and we are indebted to the MAD team of Paola Amico, Enrico Marchetti
and Johann Kolb (with particular thanks to Johann for the Strehl maps).
MAC acknowledges financial support from the Science and Technology
Facilities Council (STFC). We thank the referee, Morten Andersen,
and Nolan Walborn and Jesus Ma\'{i}z Apell\'{a}niz for their suggestions
and comments which improved this article. We also thank the HAWK-I
instrument and commissioning teams for making
the 30~Dor frames available and John Pritchard for his advice with
those images, Robert Gruendl for correspondence regarding the YSO
catalogue, Yazan Al Momany for a fateful conversation over coffee in
Santiago and subsequent advice on the individual frames method, and
Peter Stetson for making his routines available to us.
\vspace{-0.2in}
|
\section{Introduction}
Quantum mechanics together with general relativity predicts that
black hole behaves like a black body, emitting thermal radiations,
with a temperature proportional to its surface gravity at the black
hole horizon and with an entropy proportional to its horizon
area~\cite{Bek,Haw}. The Hawking temperature and horizon entropy
together with the black hole mass obey the first law of black hole
thermodynamics~\cite{BCH}. Since these seminal works in the 1970s, the
relation between thermodynamics and spacetime horizons has been
widely discussed, and further developments can be found in a nice
review~\cite{Wald:1999vt}.
The study on the relation between thermodynamics and gravity theory
can be classified into two categories: One is to discuss
thermodynamics associated with spacetime horizons in the Einstein
general relativity or in generalized gravity theories. The study of
stationary black hole thermodynamics belongs to this category.
Recently the discussions on thermodynamic properties associated with
event horizon of stationary black holes have been generalized to
various horizons of dynamical spacetimes~\cite{Ashtekar:2004cn}. For
example, it has been shown that there also exists Hawking radiation
associated with an apparent horizon of a Friedmann-Robertson-Walker (FRW)
universe~\cite{CCH}. The other is more interesting: to derive equations
of motion of the gravitational field from thermodynamics. In 1995, Jacobson
derived the Einstein equation by employing the fundamental Clausius
relation $\delta Q =TdS$ together with the equivalence
principle~\cite{Jacob}. Here the key idea is to demand that this
relation should hold for all the local Rindler causal horizons
through each spacetime point, with $\delta Q$ and $T$ interpreted as
the energy flux and Unruh temperature seen by an accelerated
observer just inside the horizon. The entropy $S$ is assumed to be
proportional to the area of the Rindler horizon. In this way, the
Einstein equation is nothing but an equation of state of spacetime.
In addition, assuming the apparent horizon of a
FRW universe has temperature $T$ and
entropy $S$ satisfying $T = 1/(2\pi\tilde r_A)$ and $S = A/(4G)$,
where $\tilde r_A$ is the radius of the apparent horizon and $A$ is
the area of the apparent horizon, Cai and Kim~\cite{CK} are able to
derive Friedmann equations of the FRW universe with any spatial
curvature by applying the Clausius relation to the apparent horizon of
the FRW universe. This approach also holds for Gauss-Bonnet gravity
and the more general Lovelock gravity. For more discussions on the
relation between the first law of thermodynamics and Friedmann
equations in diverse gravity theories, see~\cite{CCH08,Cai07} and
references therein. In the black hole spacetimes, the relation
between the first law of thermodynamics and gravitational field
equations has also been studied~\cite{Pad02}. For a recent review on
this topic and some relevant issues, see \cite{Pad09}.
In a recent paper by Verlinde~\cite{Verl}, with the holographic
assumption, gravity is explained as an entropic force caused by
changes in the information associated with the positions of material
bodies. Among various interesting observations made by Verlinde,
here we mention two of them. One is that with the assumption of the
entropic force together with the Unruh temperature~\cite{Unruh},
Verlinde is able to derive the second law of Newton. The other is
that the assumption of the entropic force together with the
holographic principle and the equipartition law of energy leads to
the Newton law of gravitation. Similar observations are also made by
Padmanabhan~\cite{Pad12}. He observed that the equipartition law of
energy for the horizon degrees of freedom combined with the
thermodynamic relation $S=E/2T$ also leads to the Newton law of
gravity. Here $S$ and $T$ are thermodynamic entropy and temperature
associated with the horizon and $E$ is the active gravitational mass
producing the gravitational acceleration in the
spacetime~\cite{Pad03}. Some very recent discussions on entropic
properties of gravity can be found
in~\cite{SG,Pad10,Cai:2010hk,Smolin:2010kk,Caravelli:2010be,Li:2010cj,
Gao:2010fw,Zhang:2010hi,Wang:2010jm,Wei:2010ww,Ling:2010zc,Zhao:2010qw,
Myung,Liu:2010na}.
On the other hand, it is well known that the Einstein general relativity
describes gravity quite well, at least classically. Therefore
the Einstein equation should imply some implications of gravity as an
entropy force. Note that various discussions made by Verlinde
are focused on the Newtonian gravity, namely in the nonrelativistic
case. Therefore it is quite interesting and important to see how
gravity appears as an entropic force in the relativistic gravity
theory. In this paper we will focus on the Einstein theory of gravity,
namely general relativity.
Note that the entropy force for a system (with many degrees of
freedom) is a macroscopic force, and it is induced by the
statistical tendency to increase the entropy of the system. So a
natural starting point to consider in the Einstein general relativity is
causal horizon of spacetimes because there exist well understood
temperature and entropy associated with the causal horizon. In this
paper, we will mainly focus on the trapping horizon of a general
spherically symmetric dynamical spacetime and explore how the Einstein
equation shows as an entropy-force-like equation. We will also
discuss the case away from the trapping horizon.
This paper is organized as follows: In Sec.~II, starting from
the Einstein equation, we show thermodynamics associated with the trapping
horizon in a general spherically symmetric dynamical spacetime. In
Sec.~III, we define a gravitational potential by employing the Kodama
vector and generalize Verlinde's argument to dynamical
spacetimes, which relates the gravitational potential to the
surface gravity of the trapping horizon. In Sec.~IV, we discuss how
the gravity on the trapping horizon appears as an entropy force. In
Sec.~V, we use a Smarr-like formula and the holographic assumption
of horizon entropy to derive the Newton gravity. In Sec.~VI, following
Verlinde~\cite{Verl} and Smolin~\cite{Smolin:2010kk}, we give some
discussions on the entropy force from the point of view of quantum
fluctuation. Section~VII is devoted to the conclusion and discussion.
\section{General spherically symmetric spacetime}
Let us consider a general spherically symmetric spacetime
$(\mathcal{M}^n,g_{\mu\nu})$ with the metric
\begin{equation}
\label{metric} g=h_{ab}dx^adx^b+ r^2(x)d\Omega_{n-2}^2\, ,
\end{equation}
where $d\Omega_{n-2}^2$ is the line element of an $(n-2)$-sphere,
and $x^a,a=1,2$ are coordinates of the two-dimensional spacetime
which is normal to the sphere. Assume the connection of the
two-dimensional space is $D_a$ (which is associated with the
two-dimensional metric $h_{ab}$). In this spacetime the Einstein
equation can be decomposed as
\begin{equation}
\label{eqm1}
G_{ab}=-\frac{n-2}{r}D_{a}D_{b}r-\left[\frac{1}{2}(n-2)(n-3)
\left(\frac{1-D_crD^cr}{r^2}\right)-\frac{n-2}{r}D_cD^c
r\right]h_{ab}=8\pi G_n T_{ab}\, ,\\
\end{equation}
\begin{equation}
\label{eqm2}
G^{i}_j=\left[-\frac{1}{2}R^{(2)}-\frac{(n-3)(n-4)}{2}\left(\frac{1-D_crD^cr}{r^2}\right)
+\frac{n-3}{r}D_cD^c
r\right]\delta^i_j=8\pi G_n T^{i}_j\, ,
\end{equation}
where $G_n$ is the $n$-dimensional Newton constant, and
$T_{\mu\nu}=(T_{ab},T_{ij})$ is the energy-momentum tensor. The term
$R^{(2)}$ in Eq.~(\ref{eqm2}) is a scalar curvature of the
two-dimensional spacetime described by $h_{ab}$. It is obvious that
one has $T_{ab}=T_{ab}(x)$ and $T^i_j=\sigma(x)\delta^i_j$ in this
case. Substituting the relation
\begin{equation}
R^{(2)}=R+\frac{2(n-2)}{r}D_cD^cr -
(n-2)(n-3)\frac{1-D_crD^cr}{r^2}\, ,
\end{equation}
one can find that Eq.~(\ref{eqm2}) is trivially satisfied if
Eq.~(\ref{eqm1}) holds. So Eq.~(\ref{eqm1}) is the master equation
we will analyze.
The Misner-Sharp energy inside the sphere with radius $r$ is given
by~\cite{Misner:1964je}
\begin{equation}
\label{MisnerSharp} E=\frac{1}{16\pi G_n}
(n-2)\Omega_{n-2}r^{n-3}(1-D_arD^ar)\, .
\end{equation}
This is active energy inside the sphere. The properties of this
energy are discussed in some detail in
Refs.~\cite{Misner:1964je,Hayward:1994bu}. With the energy-momentum
tensor $T_{ab}$, one can define two physical quantities:
\begin{equation}
w=-\frac{1}{2}T^{a}_{a}\, ,
\end{equation}
which is called work density, and
\begin{equation}
\psi_a=T_{a}^{b}D_br+w D_ar\, ,
\end{equation}
which is called energy supply. It follows from (\ref{eqm1}) and
(\ref{eqm2}) that
\begin{equation}
\label{energydensity} w=\frac{1}{16\pi
G_n}\left[-\frac{n-2}{r}D_cD^cr
+(n-2)(n-3)\left(\frac{1-D_crD^cr}{r^2}\right)\right]\, ,
\end{equation}
and
\begin{equation}
\label{energysupply} \psi_a=\frac{1}{16\pi
G_n}\frac{n-2}{r}\left[(D_bD^br)D_ar -D_a(D^br D_br)\right]\, .
\end{equation}
Combing Eqs.~(\ref{MisnerSharp}), (\ref{energydensity}) and
(\ref{energysupply}), one has~\cite{Hayward:1997jp, Hayward:1998ee}
\begin{equation}
\label{prefirstlaw} D_aE=A\psi_a+wD_aV\, ,
\end{equation}
where $A=\Omega_{n-2}r^{n-2}$ and $V=\Omega_{n-2}r^{n-1}/(n-1)$ are
area and volume of the sphere with radius $r$, respectively. We can also
express this equation in the form $dE=A\psi+wdV$ by defining
one-form $\psi=\psi_adx^a$ and differential operator $d=dx^a D_a$.
To study causal structure of the spacetime (\ref{metric}), it is
convenient to introduce two null vector fields $\ell_a$ and $k_a$,
and write the two-dimensional metric as
$h_{ab}=-\ell_ak_b-k_a\ell_b$, where $\ell_ak^a=-1$. By calculating
the extrinsic curvature of the $(n-2)$-sphere, one gets the value of
the extrinsic curvature along the $\ell_a$ and $k_a$ directions, and
then gets the expansions of the corresponding null congruences.
These two expansions are denoted by $\theta_{(\ell)}$ and
$\theta_{(k)}$, respectively.
An $(n-2)$-dimensional sphere is called marginal if
$\theta_{(\ell)}\theta_{(k)}=0$. Similarly, an untrapped sphere is
given by $\theta_{(\ell)}\theta_{(k)}<0$, and a trapped sphere is
given by $\theta_{(\ell)}\theta_{(k)}>0$. It is found that
$\theta_{(\ell)}\theta_{(k)}\sim -D_arD^ar$. Therefore the marginal
sphere satisfies $D_arD^ar=0$. The hypersurface foliated by the
marginal spheres is called a trapping horizon. This means that
$D_arD^ar$ always vanishes on this hypersurface. Let $\xi$ be a
vector field which is tangent to the trapping horizon. We then have
\begin{equation}
\mathcal{L}_{\xi}(D_brD^br)=\xi^aD_a(D_brD^br)=0\,
,
\end{equation}
Considering Eq.~(\ref{energysupply}), on the trapping horizon,
we find
\begin{equation}
A\psi_a\xi^a= \frac{\kappa_H}{2\pi}\mathcal{L}_{\xi}S\, ,
\end{equation}
where $\mathcal{L}_{\xi}$ is a Lie derivative along $\xi$, and
\begin{equation}
\label{entropy_surfacegravity} \kappa_H=\frac{1}{2}D_aD^a r\,
,\qquad S=\frac{A_H}{4G_n}\, .
\end{equation}
The $\kappa_H$ is called surface gravity~\cite{Hayward:1997jp} and
$A_H$ is the area of the trapping horizon. By defining
$T_H=\kappa_H/2\pi$, along the vector $\xi$, we have
\begin{equation}
\mathcal{L}_{\xi}E= T_H \mathcal{L}_{\xi} S + w \mathcal{L}_{\xi}V\,
.
\end{equation}
This is the first law of the dynamical black
holes~\cite{Hayward:1997jp}. $S$ and $T$ are the Bekenstein-Hawking
entropy and Hawking temperature associated with the trapping
horizon~\cite{CHNVZ}.
The surface gravity (\ref{entropy_surfacegravity}) can also be
understood from the so-called Kodama vector field~\cite{Kodama:1979vn}:
\begin{equation}
K^a=-\epsilon^{ab}D_br\, , \qquad K_aK^a=-D_arD^ar\, .
\end{equation}
So the Kodama vector is null on the trapping horizon and timelike in
the untrapped region. In addition, on the trapping horizon one has
$K_a=D_ar$. By this vector, one can define a surface
gravity on the trapping horizon as
\begin{equation}
K^bD_{[b}K_{a]}=\kappa_H K_a\, .
\end{equation}
A straightforward calculation shows that this gives the same result
of $\kappa_H$ as (\ref{entropy_surfacegravity}). In the following
discussions, we will also use the notation $\kappa=(1/2) D_aD^ar$.
$\kappa$ reduces to the surface gravity ($\kappa_H$) on the
trapping horizon, while the physical meaning of $\kappa$ will be
shown shortly.
\section {Surface gravity and gravitational potential}
In this section, we discuss the relation between surface gravity and
gravitational potential. We note that Eq.~(\ref{energysupply}) can
be rewritten as
\begin{equation}
\label{kappapotentialgeneral}
\kappa D_a r-\frac{1}{2
}D_a(D_brD^br)=\frac{8\pi G_n }{n-2}r\psi_a\, .
\end{equation}
This equation holds not only on the trapping horizon but also in the
untrapped region. Actually, it is just a part of the Einstein
equation and is valid at each point of spacetime.
Let us first consider the vacuum case in which the energy-momentum tensor
vanishes. Then Eq.~(\ref{kappapotentialgeneral}) gives
\begin{equation}
\kappa D_ar=\frac{1}{2}D_a(D_brD^br)=\frac{1}{2}D_a(-K_bK^b)\, .
\end{equation}
Defining $e^{2\phi} \equiv -K^aK_a=D^arD_ar$, we have
\begin{equation}
\label{kappa_psivanished} \kappa D_ar=e^{2\phi}D_a\phi\, ,\qquad
\mathrm{or} \qquad \kappa=D^a rD_a\phi \, .
\end{equation}
In the static case, the Kodama vector reduces to a timelike Killing
vector, and $\phi$ is the generalized Newton potential in general
relativity. Here, by using the Kodama vector, we have generalized to
the dynamical spacetime from the static one discussed by
Verlinde~\cite{Verl}, where a timelike Killing vector is employed to
relate the Newton potential to gravitational acceleration.
Let $n_a$ be an arbitrary vector field. We then have
\begin{equation}
\label{kappapotential}
\kappa \mathcal{L}_n
r=e^{2\phi}\mathcal{L}_n\phi\, ,\qquad
\left(\frac{\kappa}{2\pi}\right)\mathcal{L}_n
S=An^a\left[\frac{n-2}{8\pi
G_n}\left(\frac{e^{2\phi}}{r}\right)D_a\phi\right]\, ,
\end{equation}
where $S$ is given by $A/(4G_n)$.
Next, to discuss the general case with matter fields, let us assume
that the metric $h_{ab}$ and the energy-momentum tensor $T_{ab}$ can
be written as
\begin{equation}
h_{ab}=-u_au_b + v_av_b\, ,
\end{equation}
and
\begin{equation}
T_{ab}=\alpha u_au_b + \beta (u_av_b + v_a u_b)+ \gamma v_av_b\, ,
\end{equation}
where $u_au^a=-1$, $v_av^a=1$ and $u_av^a=0$. The quantities
$\alpha, \beta$ and $\gamma$ are functions of the two-dimensional
coordinates $x^a$. In that case we have
\begin{equation}
w=\frac{1}{2}(\alpha - \gamma)\, ,
\end{equation}
\begin{equation}
\label{energysupply1} \psi_a u^a= -\frac{1}{2}(\alpha +
\gamma)\mathcal{L}_u r - \beta \mathcal{L}_{v}r\, ,
\end{equation}
\begin{equation}
\label{energysupply2} \psi_a v^a= \frac{1}{2}(\alpha +
\gamma)\mathcal{L}_v r + \beta \mathcal{L}_{u}r\, .
\end{equation}
Further we can obtain
\begin{equation}
\kappa \mathcal{L}_u r=e^{2\phi}\mathcal{L}_u\phi - \frac{4\pi
G_n}{n-2}r\left[(\alpha + \gamma)\mathcal{L}_u r + 2\beta
\mathcal{L}_{v}r\right]\, ,
\end{equation}
\begin{equation}
\kappa \mathcal{L}_v r=e^{2\phi}\mathcal{L}_v\phi + \frac{4\pi
G_n}{n-2}r\left[(\alpha + \gamma)\mathcal{L}_v r + 2\beta
\mathcal{L}_{u}r\right]\, ,
\end{equation}
and then
\begin{equation}
\label{LuvE} \mathcal{L}_u E= -\gamma \mathcal{L}_uV-\beta
\mathcal{L}_vV\, ,\qquad \mathcal{L}_v E= \alpha
\mathcal{L}_vV+\beta \mathcal{L}_uV\, .
\end{equation}
Thus, when the energy-momentum tensor does not vanish, the relation
between the surface gravity $\kappa$ and gravitational potential in
Eq.~(\ref{kappapotential}) has to be modified. One has to consider
the contribution of the matter fields.
In a general case, $\psi_a$'s do not vanish. In some special cases,
for example Reissner-Nordstr\"om spacetime, however, one has vanishing
$\psi_a$ with a nonvanishing $w$ (see the Appendix). Note that for a
FRW universe with $\beta=0$ and $\alpha+\gamma=0$, one has also
$\psi_a=0$. In those special cases, (\ref{kappapotential}) still
holds, although matter fields are not absent.
\section{gravity as entropy force}
On the trapping horizon, Eq.~(\ref{kappapotential}) implies some
relation between the change of entropy and the gravitational
potential. For an arbitrary vector field $n$, from
Eq.~(\ref{kappapotentialgeneral}), we find
\begin{equation}
\mathcal{L}_n
S=(\varphi_g +\varphi_m)A\, ,
\end{equation}
where
\begin{equation}
\varphi_g =s_g^a n_a\, ,\qquad \varphi_m=s_m^a n_a\, .
\end{equation}
Here $\varphi_m$ is the value of entropy flux $s^a_m$ induced by
the matter field along the $n$ direction~\cite{Hayward:2004dv,
Hayward:2004fz}, and $s_m^a$ is defined as
\begin{equation}\left(\frac{\kappa}{2\pi}\right) s_m^a=\psi^a\, .
\end{equation}
Similarly $\varphi_g$ can be understood as the entropy flux
$s_g^a$ given by the change of gravitational potential along the $n$
direction. $s_g^a$ is defined by
\begin{equation}
\left(\frac{\kappa}{2\pi}\right)s_g^a=\left(\frac{n-2}{8\pi
G_n}\right)\left(\frac{e^{2\phi}}{r}\right)D^a\phi\, .
\end{equation}
We may understand that the term $A \varphi_g$ corresponds to the
work done by gravity. The reason is that the gravitational potential
will change along the $n$ direction for an arbitrary $n$. This
suggests that on the trapping horizon we have
\begin{equation}
\label{modifiedclausiurelation} T_H\mathcal{L}_n S= n^a F_a +
\delta_n Q\, ,
\end{equation}
where $\delta_n Q=A\psi_an^a$. To understand the meaning of $F_a$ in
this equation, let us define
\begin{equation}
U_a=e^{2\phi}D_a\phi\, .
\end{equation}
Obviously, this $U_a$ has a dimension of gravitational acceleration.
Note that on the trapping horizon, we have
\begin{equation}
E=\frac{1}{16\pi G_n} (n-2)\Omega_{n-2}r^{n-3}\, ,
\end{equation}
which leads to
\begin{equation}
\label{FEU} n^aF_a=n^a(2EU_a)\, .
\end{equation}
This suggests that $F_a$ is a force --- the force acting on the
active energy inside the marginal sphere.
Equation (\ref{modifiedclausiurelation}) is valid on the trapping horizon
only. The term $\delta_n Q =A\psi_an^a$ is nothing, but heat flux
caused by the matter fields.
Here some remarks are in order.
\begin{enumerate}
\item When the vector field $n$ is tangent to a surface with a fixed
gravity potential (equipotential surface, the trapping horizon is a
kind of equipotential surface), the force along the $n$ direction
does not exist. In this case, under the Lie derivative
$\mathcal{L}_n$, the marginal sphere changes to another marginal
sphere (of course inside the trapping horizon). The modified
Clausius relation (\ref{modifiedclausiurelation}) becomes normal
one, i.e., $T_H\mathcal{L}_nS=\delta_n Q$.
\item If $n$ has a component which is normal to the trapping horizon,
the marginal sphere tends to change to an untrapped sphere. There is
a change of gravitational potential along the $n$ direction. In this
case, the work term of gravity appears. In other words the force
$F_a$ is present in this case. It is clear from
(\ref{modifiedclausiurelation}) that the force appears when the
entropy associated with the trapping horizon changes. In this way
the force can be understood as an entropy force in the spirit of
arguments by Verlinde.
\item At the moment, it is not clear whether it is valid that the
force acts as an entropy force on the untrapped sphere because in
that case it is not clear whether the surface gravity $\kappa$ and
$A/(4G_n)$ have the interpretation as temperature and entropy for an
untrapped sphere. On this point we will have more discussions below.
\item Combing Eqs.~(\ref{prefirstlaw}) and
(\ref{modifiedclausiurelation}), we have
\begin{equation}
\label{modifiedfirstlaw} \mathcal{L}_n E= T_H\mathcal{L}_n S +w
\mathcal{L}_n V - n^a F_a\, .
\end{equation}
This equation is a consequence of the Einstein equation on the
trapping horizon. Along an arbitrary vector field $n$, we should
consider not only the work done by the matter fields, i.e.,
$w\mathcal{L}_nV$, but also the work made by gravity, which is just
the term $n^aF_a$.
\item Since the surface gravity can be expressed as
\begin{equation}
\label{kappabar} \frac{\kappa}{2\pi}=\frac{1}{4\pi}D_aD^ar=\frac{4
G_n}{n-2}\left[\left(\frac{n-3}{\Omega_{n-2}}\right)\left(\frac{E}{r^{n-2}}\right)-rw\right]\,
,
\end{equation}
we may define a new surface gravity $\bar\kappa$ as
\begin{equation}
\label{kappabar1} \bar{\kappa}= \kappa +\frac{8\pi G_n}{n-2}rw =
\frac{8\pi
G_n}{\Omega_{n-2}}\left(\frac{n-3}{n-2}\right)\left(\frac{E}{r^{n-2}}\right)
\, .
\end{equation}
On the tapping horizon, this new surface gravity is just the
so-called ``effective surface gravity" proposed by Ashtekar
{\it et. al.}~\cite{Ashtekar:2004cn}. From the definition (\ref{kappabar1}),
this effective surface gravity reduces to the Newton surface gravity if
we take nonrelativistic limit of $E$ (that is, replacing the energy
$E$ by mass $M$ times $c^2$). Thus, the term $w \mathcal{L}_n V$ in
Eq.~(\ref{modifiedfirstlaw}) can be absorbed into $T_H\mathcal{L}_n
S$ to give $\bar{T}_H\mathcal{L}_n S$ with definition
$$\bar{T}_H=\frac{\bar{\kappa}_H}{2\pi}=\frac{(n-3)}{4\pi r_H}\, .$$
The first law (\ref{modifiedfirstlaw}) then becomes
\begin{equation}
\label{kappabarfirstlaw} \mathcal{L}_n E= \bar{T}_H\mathcal{L}_n S -
n^a F_a\, .
\end{equation}
Along the trapping horizon, the force disappears and this equation
changes to
\begin{equation}
\mathcal{L}_n E= \bar{T}_H\mathcal{L}_n S \, .
\end{equation}
Everything becomes simple with this effective surface gravity
$\bar{\kappa}_H$. Although this effective surface gravity does not
reduce in the static limit to the standard surface gravity of static
black holes, for example, Reissner-Nordstr\"om black holes (see the
Appendix), it is enlightening when studying dynamical spacetimes. With
this effective surface gravity, one immediately has $\mathcal{L}_n
E=\delta_nQ$, and the work term $w\mathcal{L}_n V$ disappears.
Unfortunately, for dynamical black holes, the definitions of the
surface gravity are far from clear so far~\cite{Nielsen:2007ac}.
Namely it is still not very clear which definition of surface
gravity is indeed related to Hawking temperature associated with
trapping horizon.
\item From Eqs.~(\ref{modifiedclausiurelation}) and (\ref{FEU}),
when $\psi_a$ vanish, we have
\begin{equation}
\label{psivanished}
T_H\mathcal{L}_n S= n^aF_a=n^a(2EU_a)\, .
\end{equation}
This clearly indicates that the gravity comes from the entropy
force: gravity appears as a change of entropy. However, when the
energy support $\psi_a$ do not vanish, one has to consider the
contribution of the heat flux $\delta_n Q$, which also causes the
change of entropy. In addition, on the trapping horizon, if the
Kodama vector is a Killing vector, the gravity indeed appears as an
entropy force.
In a general case when matter fields are present, on the trapping
horizon, we have
\begin{equation}
\label{entropyforcegeneral} T_H\mathcal{L}_n S=
n^aF_a+\delta_nQ=n^a(2EU_a)+\delta_n Q=2\kappa_H E \mathcal{L}_nr \,
.
\end{equation}
It should be stressed here that we have
\begin{equation}
\label{entropyforce1} T_H\mathcal{L}_n S=2\kappa_H E
\mathcal{L}_nr\, ,
\end{equation}
even when $\psi_a$ do not vanish. In this case, however, the term
$2\kappa_H E\mathcal{L}_n r$ includes the contributions from the
heat flux given by the matter fields and the work done by
gravity. Therefore only when $\psi_a$ vanish, $2\kappa_H
E\mathcal{L}_n r$ stands for the work done by gravity.
\item To further understand the gravitational acceleration $U_a$, let
us consider the case without matter. This means $\psi_a=w=0$. In
this case, $E$ is a constant (this can be seen from
Eq.~(\ref{prefirstlaw}) or (\ref{LuvE})). The gravitational
acceleration $U_a$ can be expressed as
\begin{equation}
U_a=e^{2\phi}D_a\phi=\frac{1}{2}D_a(D_brD^br)=-\frac{8\pi
G_n}{(n-2)\Omega_{n-2}}D_a\left(\frac{E}{r^{n-3}}\right)\, .
\end{equation}
Considering $E$ is a constant, we have
\begin{equation}
\label{gravitationlacceleration}
U_a=\left(\frac{n-3}{n-2}\right)\frac{8\pi
G_n}{\Omega_{n-2}}\left(\frac{E}{r^{n-2}}\right)D_ar\, .
\end{equation}
This is the gravitational acceleration on the trapping horizon
provided by the energy $E$. In the nonrelativistic limit, $E$
reduces to the Newton mass $M$ (the light velocity is set to be
unity). So $U_a$ indeed gives us the correct gravitational
acceleration. Note that this calculation is also valid in the
untrapped region.
On the trapping horizon, we have
\begin{equation}
n^aF_a= n^a(2 E U_a)=2 \left(\frac{n-3}{n-2}\right)\left(\frac{8\pi
G_n}{\Omega_{n-2}}\right)\left(\frac{E^2}{r^{n-2}}\right)\mathcal{L}_nr\,
.
\end{equation}
Remarkably, this force has the form of the Newton gravity if we take
the nonrelativistic limit where $E$ is replaced by mass $M$. But
there is an additional factor $``2"$ in the second and last terms,
compared to the standard form of the Newton gravity. The factor
$``2"$ might come from the self-gravitating effect here since the
Newton force appears as a probe approximation.
\end{enumerate}
{}From the above discussions, we can conclude that {\it gravity
indeed can be viewed as an entropy force on the trapping horizon}; it
is particularly clear when the energy supply $\psi_a$ is absent on
the trapping horizon. This conclusion is based on the definition of
the quasilocal energy $E$ in Eq.~(\ref{MisnerSharp}) and the
temperature $T_H$ in Eq.~(\ref{entropy_surfacegravity}).
Furthermore, if one uses the effective surface gravity
$\bar{\kappa}_H$ and the corresponding temperature $\bar{T}_H$ when
matter fields are present, gravity can be viewed as an entropy force
if the variation of the energy $\mathcal{L}_nE$ vanishes on the
trapping horizon.
\section{Smarr-like equations and Holographic Assumption of Entropy}
It is interesting to note that the relation among the
thermodynamical quantities discussed in the previous sections can be
put in a simple form. On the trapping horizon, we find from the
expression of the surface gravity~(\ref{kappabar}) that
\begin{equation}
\frac{\kappa_H}{2\pi}=\frac{4
G_n}{n-2}\left[\left(\frac{n-3}{\Omega_{n-2}}\right)
\left(\frac{E}{r^{n-2}}\right)-rw\right]\,
.
\end{equation}
A straightforward calculation gives
\begin{equation}
\left(\frac{\kappa_H}{2\pi}\right)\left(\frac{A}{4G_n}\right)
=\left(\frac{n-3}{n-2}\right)E-\left(\frac{n-1}{n-2}\right)wV\,
.
\end{equation}
Identifying $T_H=\kappa_H/2\pi$ and $S=A_H/4G_n$, we get
\begin{equation}
\label{smarr}
(n-2)T_HS=(n-3)E-(n-1)wV\, .
\end{equation}
Further if we use the effective surface gravity $\bar{\kappa}$
instead of $\kappa$, Eq.~(\ref{smarr}) changes to
\begin{equation}
\label{TSEkappabar} (n-2)\bar{T}_HS=(n-3)E\, .
\end{equation}
Equations (\ref{smarr}) and (\ref{TSEkappabar}) are very similar to
the Smarr formula in general relativity, and we call them Smarr-like
equations. Note here that they take the quasilocal form. These
relations among energy, temperature and entropy give us a lot of
implications. For instance, since the entropy is determined by the
area of the marginal sphere, it means that the physical degrees of
freedom are determined by the marginal surface. One can imagine that
there are some bits living on the marginal sphere which give the
same amount of the entropy. If we further assume there are no
interactions among these bits, from statistic physics, at least at
high temperature, we can use the equipartition of energy to link the
energy $E$ and $T_H$. This idea is used to investigate gravity as
an entropy force by Padmanabhan~\cite{Pad12} and Verlinde~\cite{Verl}.
Now let us consider the case without matter with $w=0$. Assume there
are $N$ bits associated with the marginal surface with~\cite{Verl}
\begin{equation}
\label{Ndefinition}
N=\frac{1}{2}\left(\frac{n-2}{n-3}\right)\left(\frac{A}{G_n}\right)\,
.
\end{equation}
The relation of the entropy and $N$ is given by
\begin{equation}
\label{SNRelation}
S=\frac{1}{2}\left(\frac{n-3}{n-2}\right)Nk_B\, ,
\end{equation}
where the Boltzmann constant $k_B$ is recovered. Because $N$ bits
have the energy $(1/2) N k_B T_H$, this gives
\begin{equation}
\frac{1}{2}N k_B T_H=\left(\frac{n-2}{n-3}\right)T_H S=E\, .
\end{equation}
Thus when the matter fields are absent, the assumption of the
equipartition of energy is consistent with the Smarr-like equation.
In the presence of matter, the law of the equipartition of energy is
broken by the term including $wV$. However, since the Smarr-like
equation (\ref{smarr}) is always satisfied, with the holographic
assumption that the entropy is given by (\ref{SNRelation}), we still have
\begin{equation}
\label{holographicentropy} \frac{1}{2}N k_B
T_H=\left(\frac{n-2}{n-3}\right)T_H
S=E-\left(\frac{n-1}{n-3}\right)wV\, .
\end{equation}
It follows from Eq.~(\ref{TSEkappabar}) that the equipartition of
energy can be always used if we use the effective surface gravity
$\bar{\kappa}_H$ and the corresponding temperature $\bar{T}_H$.
It is interesting to study the entropy force by using this
description of the physical degrees of freedom. Can we get the
relation (\ref{entropyforce1}) just from the above holographic
scenario? We can imagine that there are $N$ bits living on the
marginal surface. Every bit has energy $\frac{1}{2} k_B T_H$. So
along the direction normal to the trapping horizon, it will feel
force $\frac{1}{2} \kappa_H k_B T_H$, and the total force is given
by
\begin{equation}
\label{confusion} N\left(\frac{1}{2} \kappa_H k_B
T_H\right)=\kappa_H E\ne 2 \kappa_H E\, .
\end{equation}
However, it is different from the result~(\ref{entropyforce1}) by a
factor 2. In fact, it is expected because the simple counting
does not include the self-gravitating effect and the result
(\ref{confusion}) appears in the probe approximation.
\section{entropy force from quantum fluctuation}
\label{quantum}
When $\psi_a=0$, it can be clearly seen that the gravitational
force appears as entropy force (\ref{psivanished}) in the Einstein
general relativity. However, it should be noted that this is just a
consequence of the Einstein equation together with thermodynamic
properties of the horizon. The change of the entropy in
(\ref{psivanished}) actually comes from the change of the area of
the marginal sphere. So the variation of the entropy is geometric,
and has no direct relation to quantum behavior of the black hole
horizon. However, it is clear that there must exist underlying
microscopic degrees of freedom associated with black hole horizon
entropy.
We now apply a similar discussion made by Verlinde~\cite{Verl} (the
so-called thought experiment of Bekenstein) to the trapping horizon
of the dynamical spacetime. Consider a test particle (with rest
mass $m$) staying just outside the trapping horizon with some
distance, $\Delta x$, from the trapping horizon. As a test particle,
we assume that it will not change the background geometry. However,
if the value of the distance $\Delta x$ is about the Compton
wavelength of the particle, this particle should be viewed as a part
of the dynamical black hole. Then the entropy of the black hole has
a fluctuation $\Delta S$ even though the geometry of the trapping
horizon is not supposed to change. Here let us concentrate only on
the four-dimensional case, with generalization to other dimensions being
straightforward. The change of the entropy is given by
\begin{equation}
\label{changeentropy}
\Delta S =2\pi k_B \frac{mc}{\hbar}\Delta x\,
.
\end{equation}
The temperature of the system is given by
\begin{equation}
k_B T_H= 2 \left(\frac{E}{N}\right)\, .
\end{equation}
This is correct only for the case without matter field.
Note that the number $N$ is given by
\begin{equation}
N=\frac{A_Hc^3}{G_4 \hbar}\, .
\end{equation}
Substituting the expression of $N$ and $T_H$ into Eq.~(\ref{changeentropy}),
we have
\begin{equation}
T_H\Delta S=G_4 \frac{ m E/c^2}{r^2 }\Delta x\, .
\end{equation}
Thus, one can identify the right hand side of the above equation as
the work done by some entropy force. It is clear that this force is
similar to the Newton force
\begin{equation}
f=G_4 \frac{ m E/c^2}{r^2 }\, .
\end{equation}
We emphasize that our discussion is restricted to the region near
the trapping horizon because $T_H$ is the Hawking temperature of
the dynamical black hole.
Note that in the case without matter, the relation
$2(\kappa/2\pi)S=E$ is also valid in the untrapped region. If we
assume that there is a ``temperature" $T$ associated with $\kappa$
and the entropy is $S=A/4G$, by using the similar reasoning, we find
the corresponding gravitational acceleration is given by
$\kappa/c^2$ which is similar to (4.18):
\begin{equation}
G_4 \frac{ E/c^2}{r^2 }.
\end{equation}
This is just the Newton gravitational acceleration provided by the
active energy $E$ inside the untrapped sphere with radius $r$. We
will give more discussions on this ``temperature" $T$ later.
Let us now turn to the case with matter. In this case, the
equipartition law of energy is violated, but the Smarr-like formula
together with the holographic assumption of the entropy can be used.
For the case of $n=4$, Eq.~(\ref{holographicentropy}) becomes
\begin{equation}
2 T_H S= \frac{1}{2}Nk_B T_H= E-3 wV\, .
\end{equation}
By using a similar discussion, we arrive at
\begin{equation}
f=G_4 \frac{ m (E-3wV)/c^2}{r^2 }=G_4 m \left(\frac{E}{r^2}-4\pi r
w\right)/c^2=m \kappa_H /c^2\, .
\end{equation}
where we have used $V=4\pi r^3/3$. So the force is nothing but the
surface gravity $\kappa_H $ times the rest mass $m$. Note that we
have used the unit of $c=1$ in the discussion of previous sections.
It should be noted that we have not used any nonrelativistic limit
till now.
Our conclusion (in unit of $c=1$) is that near the trapping horizon,
for the test particle with rest mass $m$, the entropy force given by
Verlinde is just the Newtonian force $\kappa_H m/c^2$ with
gravitational acceleration $\kappa_H/c^2$.
We can understand the entropy force from another point of view which
is similar to the discussion by Smolin~\cite{Smolin:2010kk}: Suppose
that there is a particle with mass $m$ ``pulled away" from the
horizon. This pulling away should be understood as quantum
fluctuations. Assume that the distance of the particle from horizon
is given by $\Delta x$. This $\Delta x$ should be within the Compton
wavelength of the particle, so the particle still belong to the
horizon. So the passive energy of this system will not change, and
the change of the active energy $E$ is then just given by
\begin{equation}
\Delta E= F\Delta x=T_H \Delta S\, .
\end{equation}
Assume the change of the entropy is still given by Eq.~(\ref{changeentropy}).
Then similar discussions can lead to
\begin{equation}
F=G_4 \frac{ (E-3wV)/c^2}{r^2 }m=m\kappa_H/c^2\, .
\end{equation}
Assume that the mass $m$ carries $q$ ($q<< N$) bits of information,
and we then have
\begin{equation}
mc^2=\frac{1}{2}q k_B T_H \, ,
\end{equation}
and we obtain the force per bit as
\begin{equation}
f=\frac{F}{q}=\left(\frac{1}{2}k_B T_H\right)\kappa_H/c^4\, .
\end{equation}
Namely one bit information on the horizon will feel a force
$(\frac{1}{2}k_BT_H)\kappa_H/c^4$. As we have mentioned at the end
of Sec.~V, there is a factor 2 difference between this kind of probe
approximation and our result~(\ref{entropyforce1}). In this
holographic description, the procedure to get
results~(\ref{entropyforce1}) should be understood as: extract all
bits on the marginal sphere to a very nearby untrapped sphere.
Obviously, the probe (or test particle) approximation is not valid
in this case.
\section{Conclusion and Discussion}
In this paper we mainly investigated the issue that gravity appears
as an entropy force from Einstein's equation in a general
spherically symmetric spacetime. On the trapping horizon of the
spacetime, we found that the gravity acting on the marginal sphere
indeed can be identified as an entropy force when the energy supply
$\psi_a$ vanishes on the trapping horizon. We noticed that when matter
fields are present, the heat flux also leads to the change of
entropy. With the holographic assumption of the entropy, following
Verlinde and Smolin, we showed that the entropy force induced by
quantum fluctuations is measured by the surface gravity on the
trapping horizon.
We give further remarks here.
In most parts of the present paper, we have focused on the trapping
horizon, because on the trapping horizon, we have well-defined
temperature and entropy associated with the trapping horizon. Off
the trapping horizon, although the surface gravity (\ref{kappabar})
on the untrapped sphere has a dimension of gravitational
acceleration, its physical meaning is still not very clear. In the
static case, the temperature at a fixed $r$ surface outside the
horizon is given by the Tolman redshift relation $e^{-\phi}T_H$ with
a redshift factor $e^{-\phi}$, provided the system is in a
thermal equilibrium. This is the temperature measured by a static
observer at the constant $r$ surface. Obviously, it is not equal to
$T=\kappa/2\pi$, while in the case without matter, this $\kappa$
indeed reduces in the nonrelativistic limit to the Newton gravity
acceleration at the surface with radius $r$. Therefore the
``temperature" $T=\kappa/2\pi$ can be understood as a local Unruh
temperature given by a Rindler observer with acceleration $\kappa$.
Indeed, as it is well known that some curved spacetimes like a
Schwarzschild spacetime can be embedded in a higher-dimensional flat
spacetime, a static observer at the constant $r$ surface is mapped
into a Rindler observer in the higher-dimensional spacetime. The
surface gravity $\kappa$ is just the Unruh temperature measured by
the Rindler observer in the higher-dimensional spacetime. Once
accepting $T$ as a real temperature in the above sense, we have from
Einstein's equation that
\begin{equation}
\label{smarrlibaration} (n-2)TS=(n-3)E-(n-1)wV\, ,
\end{equation}
which reduces to Eq.~(\ref{smarr}) on the trapping horizon. Here now
$E$ is just the Misner-Sharp energy given by (\ref{MisnerSharp}),
and $S=A/4G_n$ for the sphere with an arbitrary radius $r$.
Employing (\ref{smarrlibaration}) with the holographic assumption of
$S$, one can find the force $m\kappa/c^2$ acting on a test particle
with mass $m$. From the expression (\ref{kappabar}), we find that
this is just the Newton gravity produced by $E$ when the matter
fields are absent.
Ashtekar's ``effective surface gravity" $\bar{\kappa}$
(\ref{kappabar}) is simple, which has a similar form to the Newton
gravity acceleration if we take the nonrelativistic limit of $E$.
With this definition of surface gravity, the corresponding
temperature $\bar{T}_H$, entropy and energy satisfy the relation
(\ref{TSEkappabar}). In this case since we have (\ref{TSEkappabar}),
the equipartition law of energy holds and on the trapping horizon,
we can obtain
\begin{equation}
\label{7eq2} m\bar{\kappa}_H/c^2=G_4 \frac{ m E/c^2}{r^2 }\, .
\end{equation}
This is just the Newton gravity produced by the active energy $E$.
With the same argument as the above, defining the temperature $\bar
T=\bar \kappa/2\pi$ on the untrapped surface, and still employing
the relation (\ref{TSEkappabar}), we can obtain the Newton gravity
(\ref{7eq2}) on an arbitrary sphere with radius $r$.
In addition, in the present paper, we have focused on the future
outer trapping horizon of spacetime, and there the surface gravity
$\kappa_H>0$. In fact, our discussion is also valid on the inner
trapping horizon, where surface gravity $\kappa_H$ is negative. The
apparent horizon of the FRW universe is just this case~\cite{CK,CCH}. In
this case, the temperature is defined as $T_H=|\kappa_H|/2\pi$, and
Eq.~(\ref{entropyforcegeneral}) then becomes
\begin{equation}
T_H\mathcal{L}_n S=n^a(2|\kappa_H| E U_a)-\delta_n Q\, .
\end{equation}
The entropy force also appears naturally. This is particularly clear
in the case of the de-Sitter spacetime, in which case $\psi_a=0$ and
$\delta_n Q=0$, and gravity appears as an entropy force.
\section*{Acknowledgements}
RGC thanks Y.S. Myung for useful discussions and he is supported
partially by grants from NSFC, China (No.10821504 and No.10975168)
and the Ministry of Science and Technology of China national basic
research Program (973 Program) under grant No. 2010CB833004. LMC and
NO were supported in part by the Grants-in-Aid for Scientific
Research Fund of the JSPS No. 20540283 and No. 21$\cdot$\,09225, and
also by the Japan-U.K. Research Cooperative Program.
|
\section{Matching of the phenomenological SU(3) and SU(4)
$PBB$ and $VBB$ interaction Lagrangians}
First we consider the $PBB$ interaction that is the coupling of a
pseudoscalar ($P$) meson to two baryons ($BB$).
In flavor SU(3) the couplings are generated
from the ${\cal O}(p)$ term of
chiral perturbation theory (ChPT)~\cite{ChPT} describing
the coupling of baryon fields with the chiral fields:
\begin{eqnarray}
{\cal L}_{PBB}^{SU_3} = -\frac{D}{F_P\sqrt{2}} (m_B + m_{\bar B})
{\rm tr} \Big(\bar B i \gamma^5 \{ P B \}\Big)
\, - \, \frac{F}{F_P\sqrt{2}} (m_B + m_{\bar B}) {\rm tr}
\Big(\bar B i \gamma^5 [ P B ]\Big) \, .
\end{eqnarray}
Here $F_P = F_\pi = 92.4$ MeV is the
leptonic decay constant; $D$ and $F$ are the baryon axial coupling constants
(we restrict to the SU(3) symmetric limit, where $D = 3F/2 = 3g_A/5$
with $g_A = 1.2695$ being the nucleon axial charge)
the symbols ${\rm tr}$, $\{ \ldots \}$ and $[ \ldots ]$ denote the trace
over flavor matrices, anticommutator and commutator, respectively.
We replace the pseudovector coupling by the pseudoscalar one
considering on-mass-shell baryons.
The SU(3) baryon $B$ and pseudoscalar meson $P$
matrices read as:
\begin{eqnarray}
B =
\left(
\begin{array}{ccc}
\Sigma^0/\sqrt{2} + \Lambda/\sqrt{6}\,\, & \,\, \Sigma^+ \,\, & \, p \\
\Sigma^- \,\, & \,\, -\Sigma^0/\sqrt{2}+\Lambda/\sqrt{6}\,\, & \, n\\
\Xi^-\,\, & \,\, \Xi^0 \,\, & \, -2\Lambda/\sqrt{6}\\
\end{array}
\right),
\end{eqnarray}
\begin{eqnarray}
P =
\left(
\begin{array}{ccc}
\pi^0/\sqrt{2} + \eta/\sqrt{6}\,\, & \,\, \pi^+ \,\, & \, K^+ \\
\pi^- \,\, & \,\, -\pi^0/\sqrt{2}+\eta/\sqrt{6}\,\, & \, K^0\\
K^-\,\, & \,\, \bar K^0 \,\, & \, -2\eta/\sqrt{6}\\
\end{array}
\right) \, .
\end{eqnarray}
The SU(4) $PBB$ Lagrangian is given by~\cite{Okubo:1975sc}:
\begin{eqnarray}
{\cal L}_{PBB}^{SU_4} = g_1 \bar B^{kmn} i\gamma_5 P^l_k B_{lmn}
+ g_2 \bar B^{kmn} i\gamma_5 P^l_k B_{lnm} \, ,
\end{eqnarray}
where the indices $l,m,n$ of the tensor $B_{lmn}$
run from 1 to 4, representing the 20--plet of baryons
(see details in Refs.~\cite{Okubo:1975sc});
$P^l_k$ is the matrix representing the 15--plet of
pseudoscalar fields. The baryon tensor satisfies the conditions
\begin{eqnarray}
B_{lmn} + B_{mnl} + B_{nlm} = 0, \hspace*{.5cm}
B_{lmn} = B_{mln} \,.
\end{eqnarray}
The full list of physical states in terms of SU(4) tensors is
given in Ref.~\cite{Okubo:1975sc}. Here we only display a few of them:
\begin{eqnarray}
& &p = B_{112} = - 2 B_{121} = - 2 B_{211}\,, \hspace*{.25cm}
n = - B_{221} = 2 B_{212} = 2 B_{122}\,, \nonumber\\
& &\Sigma_c^{++} = B_{114} = - 2 B_{141} = - 2 B_{411}\,, \hspace*{.25cm}
\Sigma_c^0 = - B_{224} = 2 B_{242} = 2 B_{422}\,, \\
& &\pi^+ = P^2_1\,, \hspace*{.25cm}
\pi^- = P^1_2\,, \hspace*{.25cm}
D^0 = P^1_4\,, \hspace*{.25cm}
D^{\ast +} = V^2_4\,, \hspace*{.25cm}
D^{\ast 0} = V^1_4 \,. \nonumber
\end{eqnarray}
Evaluating the $\pi NN$ couplings in both versions we fix
the SU(4) couplings $g_1$ and $g_2$ as (in the SU(3) Lagrangian
we restrict to the mass degenerate case $m_B = m_{\bar B}
= m_p = 938.27$ MeV):
\begin{eqnarray}
g_{\pi NN} = g_1 - \frac{5}{4} g_2\,, \hspace*{.25cm}
g_{\pi NN} \frac{D-F}{D+F}= - \frac{g_1 + g_2}{4 \sqrt{2}}.
\end{eqnarray}
Considering the SU(3) symmetric ratio of $F$ and $D$ couplings
$F/D = 2/3$ we get
\begin{eqnarray}
g_1 = 0\,, \hspace*{.25cm}
g_2 = - \frac{4}{5} \sqrt{2} \, g_{\pi NN} \,.
\end{eqnarray}
Finally the $g_{DN\Lambda_c}$ coupling is fixed as
\begin{eqnarray}
g_{DN\Lambda_c} = - g_{\pi NN}\,.
\end{eqnarray}
In complete analogy we fix the vector meson $VBB$ couplings.
The SU(3) $VBB$ Lagrangian
can be expressed in terms of the $\rho NN$ coupling constant as:
\begin{eqnarray}
{\cal L}_{VBB}^{SU_3} = \frac{g_{\rho NN}}{\sqrt{2}}
{\rm tr} \Big(\bar B \gamma^\mu \{ V_\mu B \}\Big)
\, + \, \frac{g_{\rho NN}}{\sqrt{2}} {\rm tr}
\Big(\bar B \gamma^\mu B\Big) {\rm tr}V_\mu \,
\end{eqnarray}
where
\begin{eqnarray}
V =
\left(
\begin{array}{ccc}
\rho^0/\sqrt{2} + \omega/\sqrt{2}\,\, & \,\, \rho^+ \,\, & \, K^{\ast +} \\
\rho^- \,\, & \,\, -\rho^0/\sqrt{2}+\omega/\sqrt{2}\,\, & \, K^{\ast 0}\\
K^{\ast -}\,\, & \,\, \bar K^{\ast 0} \,\, & \, - \phi \\
\end{array}
\right).
\end{eqnarray}
The SU(4) $VBB$ Lagrangian is given by~\cite{Okubo:1975sc}:
\begin{eqnarray}
{\cal L}_{PBB}^{SU_4} = h_1 \bar B^{kmn} \gamma^\mu V^l_{\mu, k} B_{lmn}
+ h_2 \bar B^{kmn} \gamma^\mu V^l_{\mu, k} B_{lnm} \, .
\end{eqnarray}
Evaluating the $\rho NN$ couplings in both versions we fix
the SU(4) couplings $h_1$ and $h_2$ as:
\begin{eqnarray}
h_1 = 2 h_2 = \frac{8}{3\sqrt{2}} g_{\rho NN} \, .
\end{eqnarray}
Finally, the $D^\ast N\Lambda_c$ coupling is fixed as
\begin{eqnarray}
g_{D^\ast N\Lambda_c} = - \frac{\sqrt{3}}{2} \,
g_{\rho NN}\, ,
\end{eqnarray}
where for the $g_{\rho NN}$ coupling we take the SU(3) prediction of
\begin{eqnarray}
g_{\rho NN} = 6 \, .
\end{eqnarray}
|
\section{Introduction} Let $(X,x)$ be a 2-dimensional normal complex analytic germ. Let $U=X\setminus \{x\}$. Mumford (\cite{Mum}) showed the celebrated theorem
\begin{thm}[Mumford] $(X,x)$ if smooth if and only if the topological fundamental group of $U$ is trivial.
\end{thm}
This is a remarkable theorem which connects a topological notion to a
scheme-theoritic one. His theorem has been a bit refined by Flenner \cite{Fle} who showed that in fact, the conclusion remains true if one replaces the topological by the \'etale fundamental group of $U$, that is by its profinite completion. Then one can replace the analytic germ by a complete or henselian germ over an algebraically closed field $k$ of characteristic $0$.
If $k$ is an algebraically closed field $k$ of characteristic $p>0$, Mumford himself observed that the theorem is no longer true. As an example, while in characteristic $0$, the singularity $z^2+xy$ is the quotient of $\widehat{{\mathbb A}}^2$, the completion of ${\mathbb A}^2$ at the origin, by the group ${\mathbb Z}/2$ acting via ${\rm diag}(-1,-1)$, in characteristic $2$, it is the quotient of $\widehat{{\mathbb A}}^2$ by $\mu_2={\rm Spec \,} k[t]/(t^2-1)$ acting via ${\rm diag}(t,t)$. Thus $\pi^{\rm et}(U)=\pi^{\rm et}(\widehat{{\mathbb A}}^2\setminus \{0\})=0$, yet $z^2+xy$ is not smooth.
Artin asked in \cite{Ar3} whether, if $\pi^{{\rm et}}(U)$ is finite, there is always a finite morphism $ \widehat{{\mathbb A}}^2\to X$. He shows this if $(X,x)$ is a rational double point {\it loc.cit.}.
The purpose of this note is to give an answer to a similar question where one replaces the \'etale fundamental group by the Nori one. Strictly speaking, Nori
in \cite[Chapter~II]{N} defined his fundamental group-scheme for irreducible reduced schemes endowed with a rational point. But as $U$ has no rational point, one has to modify a tiny bit Nori's construction to make it work. This is done in subsection \ref{ss_loc}. While the \'etale fundamental group of $X$ is trivial, Nori's one isn't. So the right notion of Nori's fundamental group is a relative one denoted by $\pi_{{\rm loc}}(U,X,x)$ (see Lemma \ref{lem2.5}). Roughly speaking, it measures the torsors on $U$ under a finite flat $k$-group-scheme $G$ which do not come by restriction from a torsor on $X$. We show (Theorem \ref{thm4.2}) that if $\pi_{{\rm loc}}^N(U,X,x)$ is finite, then $(X,x)$ is a rational singularity, and if $\pi_{{\rm loc}}^N(U,X,x)=0$, then there is a finite morphism $f: \widehat{{\mathbb A}}^2\to X$.
This note relies on discussions the authors had during the Christmas break 2009/10 in Ivry. They have been written down by H\'el\`ene in the night when Eckart died, as a despaired sign of love.
\section{Local Nori Fundamental Groupscheme}
\subsection{Nori's construction}
Let $U$ be a scheme defined over a field $k$, endowed with a rational point $u\in U(k)$.
In \cite[Chapter~II]{N} Nori constructed the fundamental group-scheme $\pi^N(U,u)$. Let ${\mathcal C}(U,u)$ be the following category. The objects are triples $(h:V\to U, G, v)$ where $G$ is a finite $k$-group-scheme, $h$ is a $G$-principal bundle and $v\in V(k)$ with $h(v)=u$. Recall \cite[Chapter~I,2.2]{N} that a $G$-principal bundle
$h: V\to U$ is a flat morphism, together with a group action $G\times_k V\xrightarrow{\bullet} V$ such that $V\times_k G\xrightarrow{(1,\bullet)} V\times_U V$ is an isomorphism.
Then ${\rm Hom}\big((h_1:V_1\to U, G_1, v_1),(h_2:V_2\to U, G_2, v_2)\big)$ consists of the $U$-morphisms $f: V_1\to V_2$ which are compatible with the principal bundle structure.
The objects of the ind-category ${\mathcal C}^{{\rm ind}}(U,u)$ associated to ${\mathcal C}(U,u)$ are triples $(h:V\to U, G, v)$ where $G=\varprojlim_\alpha G_\alpha$ is a prosystem of finite $k$-group-schemes $G_\alpha$, $h=\varprojlim_\alpha h_\alpha, h_\alpha: V_\alpha\to U$, is a pro-$G$-principal bundle and $v=\varprojlim_\alpha v_\alpha \in Y(k)$ is a pro-point with $h(v)=u$. The morphisms are the ind-morphisms $V_1\to V_2$ over $U$ which are compatible with the principal bundle structure and such that $f(v_1)=v_2$.
Then $(U,u)$ has a fundamental group-scheme $\pi^N(U,u)$, which is then a $k$-profinite group-scheme, if by definition \cite[Chapter~II, Definition~1]{N} there is a $(\frak{h}:W\to U, \pi^N(U,u),w)\in
{\mathcal C}^{\rm ind}(U,u)$ with the property that for any $(h: V\to U, G, v)\in {\mathcal C}^{\rm ind}(U,u)$, there is a unique map
$(\frak{h}:W\to U, \pi^N(U,u),w)\to (h: V\to U, G, v)$
in ${\mathcal C}^{{\rm ind}}(U,u)$.
Nori shows \cite[Chapter~II, Lemma~1]{N} that if $G_1, G_2, G_0$ are three finite $k$-group-schemes, $h_i: V_i\to U$ are $G_i$-principal bundles, and $f_i: V_i\to V_0, i=1,2$ are principal bundle $U$-morphisms, then $V_1\times_{V_0} V_2\to Z$ is a principal bundle under $G_1\times_{G_0} G_2$, where $Z\subset U$ is a closed subscheme (no reference to the base point here). Then he shows that $(U,u)$ has a fundamental group-scheme if and only if $Z=U$ for all $(h_i: V_i\to U, G_i, y_i), f_i\in {\mathcal C}(U,u)$
and he concludes \cite[Chapter~II, Proposition~2]{N} that if $U$ is reduced and irreducible, then $(U,u)$ has a fundamental group-scheme.
\subsection{Local Nori fundamental group-scheme} \label{ss_loc}
Let $k$ be a field, let $A$ be a complete normal local $k$-algebra with maximal ideal $\frak{m}$ and residue field $k$. We define $X={\rm Spec \,} A$ and $U=X\setminus \{x\}$, where $x\in X(k)$ is the rational point associated to $\frak{m}$. So in particular, $U(k)=\emptyset$, and we have to slightly modify Nori's construction to define the group-scheme of $U$.
Let $G$ be a finite $k$-group-scheme, and let $h: V\to U$ be a $G$-principal bundle. Recall from \cite[Corollaire~6.3.2, Proposition~6.3.4]{EGA2} that the {\it integral closure} $\tilde{h}: Y\to X$ of $h$ is the {\it unique} extension $\tilde{h}: Y\to X$ of $h$ such that $Y={\rm Spec \,} B$, $B$ is the integral closure of $A$ in $j_*h_*{\mathcal O}_V$, where $j: U\to X$ is the open embedding. Then $\tilde{h}$ is finite. In particular, if $h_i: V_i\to U$ are principal bundles under the finite $k$-group-schemes $G_i$, and $f: V_1\to V_2$ is a $U$-morphism which respects the principal bundle structures, then it extends uniquely to a $X$-morphism $\tilde{f}: Y_1\to Y_2$, which is then finite.
We can now mimic Nori's construction.
\begin{defn} \label{defn2.1}
The objects of the category ${\mathcal C}_{{\rm loc}}(U,x)$ are triples $(h: V\to U, G, y)$ where $G$ is a finite $k$-group-scheme, $y\in Y(k)$ with $\tilde{h}(y)=x$, where $\tilde{h}: Y\to X$ is the integral closure of $h$ . The morphisms ${\rm Hom} \big((h_1: V_1\to U, G_1, y_1) \to (h_2: V_2\to U, G_2, y_2)\big)$ consist of $U$-morphisms $f: V_1\to V_2$ which respect the principal bundle structure and such that $\tilde{f}(y_1)=y_2$.
The objects of the ind-category ${\mathcal C}^{{\rm ind}}_{{\rm loc}}(U,x)$ associated to ${\mathcal C}_{{\rm loc}}(U,x)$ are triples $(h: V\to U, G, y)$ where $G=\varprojlim_\alpha G_\alpha$ is a pro-system of finite $k$-group-schemes,
$h=\varprojlim_\alpha h_\alpha, h_\alpha: V_\alpha\to U$, is a pro-$G$-principal bundle, and $y=\varprojlim_\alpha y_\alpha \in \varprojlim_\alpha Y_\alpha(k)$ is a pro-point in the integral closure of $V_\alpha$ mapping to $x$.
One says that $(U,x)$ has a {\it local fundamental group-scheme} $\pi^N_{{\rm loc}}(U,x)$, which is then a $k$-profinite group-scheme, if there is a $(\frak{h}:W\to U, \pi^N_{{\rm loc}}(U,x),z)\in
{\mathcal C}^{\rm ind}_{{\rm loc}}(U,x)$ with the property that for any $(h: V\to U, G, v)\in {\mathcal C}^{\rm ind}_{{\rm loc}}(U,x)$, there is a unique map
$(\frak{h}:W\to U, \pi^N_{{\rm loc}}(U,x),z)\to (h: V\to U, G, y)$
in ${\mathcal C}^{{\rm ind}}_{{\rm loc}}(U,x)$.
\end{defn}
\begin{prop} \label{prop2.3}
If $X$ is reduced and irreducible, then $(U,x)$ has a {\it local fundamental group-scheme} $\pi^N_{{\rm loc}}(U,x)$.
\end{prop}
\begin{proof}
As explained above, the condition on $X$ implies that if
$f_i: (h_i: V_i\to U, G_i, y_i)\to (h_0: V_0\to U, G_0, y_0)$ is a morphism in ${\mathcal C}_{{\rm loc}}(U,x)$, then $(V_1\times_{V_0} V_2\to U, G_1\times_{G_0} G_2, y_1\times_{y_0} y_2)\in {\mathcal C}_{{\rm loc}}(U,x)$, so
as in \cite[Chapter~II,p.87]{N}, the prosystem $\varprojlim_\alpha (h_\alpha: V_\alpha\to U, G_\alpha, y_\alpha)$ over all objects $ (h_\alpha: V_\alpha\to U, G_\alpha, y_\alpha) $ of ${\mathcal C}_{{\rm loc}}(U,x)$ is well defined. So $\pi^N_{{\rm loc}}(U,x)=\varprojlim_\alpha G_\alpha$.
\end{proof}
There is a restriction functor $\rho: {\mathcal C}(X,x)\to {\mathcal C}_{{\rm loc}}(U,x)$ which sends $(h: Y\to X, G, y)$ to its restriction $(h_U: Y\times_X U\to U, G, y)$, as the integral closure of $X$ in $Y\times_X U$ is $Y$. This defines the $k$-group-scheme homomorphism $$\rho_*: \pi^N_{{\rm loc}}(U,x)\to \pi^N(X,x).$$
\begin{prop} \label{prop2.3}
The homomorphism $\rho$ is faithfully flat.
\end{prop}
\begin{proof}
Faithful flatness of $\rho$ means that if
$(h: Y\to X, G, y)\in {\mathcal C}(X,x)$ is such that $(Y_U\to, G, y) \to (U, \{1\},x)$ factors through $(\ell: V\to U, H, y) \in {\mathcal C}_{{\rm loc}}(U,x)$, where $Y_U=Y\times_X U$, then
necessarily $(\ell: V\to U, H, y) =\rho(\ell_X: Z\to X, H, y)$ for some $(\ell_X: Z\to X, H, y) \in {\mathcal C}(X,x)$.
Let $K={\rm Ker}(G\to H)$. Since $K$ is a $k$-subgroup-scheme of $G$, it acts on $Y$. We define $Z$ to be $Y/K$. By definition, $Z_U=V$. The compositum $h: Y\to Z\to X$ is a $G$-torsor.
The embedding $Y\times_Z Y\subset Y\times_X Y$ is closed, and while restricted to $U$, it is described as $Y_U \times_k K\subset Y_U\times_k G$. Thus $Y\times_ZY$ contains the closure of $Y_U\times_k K$ in $Y\times_k G$, that is $Y\times_k K$. Thus $Y\times_k K$ consists of connected components of $Y\times_Z Y$ and moreover, if there is another connected component, it lies in $\{y\}\times_Z Y={\rm Spec \,} k$. Thus $Y\times_Z Y\cong_k Y\times_k K$ and $Y\to Z$ is a $K$-torsor. This finishes the proof.
\end{proof}
We denote by $\pi^{{\rm et}}(U,x)$ the \'etale proquotient of $\pi^N_{{\rm loc}}(U,x)$. From now on, we assume
$k=\bar k$. Then $\pi^{{\rm et}}(U,x)$ is identified with $\pi^{\rm et}(U,\eta)$ where $\eta\to U$ is a geometric generic point and $\pi^{\rm et}(U,\eta)$ is Grothendieck's \'etale fundamental group. The \'etale proquotient of $\pi^N(X,x)$ is identified with Grothendieck's fundamental group based at $x$, and is trivial by Hensel's lemma, as $A$ is complete. If $\ell$ is a prime number (including $p$), we denote by $\pi^{{\rm et, ab,}\ell}(U,x)$ the maximal pro-$\ell$-abelian quotient of $\pi^{{\rm et}}(U,x)$.
\begin{defn}
One defines $\pi^N_{\rm loc}(U,X,x)={\rm Ker}\big( \pi^N_{{\rm loc}}(U,x)\xrightarrow{\rho} \pi^N(X,x)\big)$.
\end{defn}
From the discussion, we see
\begin{lem} \label{lem2.5}
The compositum $\pi^N_{{\rm loc}}(U,X,x)\to \pi^{\rm et}(U,x)$ is surjective. In particular, if $\pi^N_{{\rm loc}}(U,X,x)$ is a finite $k$-group-scheme, $\pi^{\rm et}(U,x)$ is a finite group.
\end{lem}
\section{Construction and elementary properties of the Picard scheme for surface singularities}
Let $k$ be a field, perfect if of characteristic $p>0$, let $A$ be a complete normal local $k$-algebra with maximal ideal $\frak{m}$, $X={\rm Spec \,} A$ and $U=X\setminus \{x\}$, where $x\in X(k)$ is the rational point associated to $\frak{m}$.
In \cite[Expos\'e~XIII,Section~5]{SGA2} Grothendieck initiated the construction of a pro-system of locally algebraic $k$-group-schemes $G_n$ and a canonical isomomorphism $G(k)={\rm Pic}(U)$ with $G(k)=\varprojlim_n G_n(k)$. This construction is performed in \cite{Lip} (see overview in \cite[p.~273]{Kl}) and relies on Mumford's basic idea \cite[Section~2]{Mum} to use a desingularization of $X$, if it exists, so in characteristic $0$ or if ${\rm dim}_k X\le 2$ if $k$ has characterisistic $p>0$. We now summarize the construction and the elementary properties under the assumptions
\begin{itemize}
\item[1)] $X$ is normal
\item[2)] ${\rm dim}_k X=2.$
\end{itemize}
Let $\sigma: \tilde{X} \to X$ be a desingularization such that $\sigma^{-1}(x)_{{\rm red}}=\cup_i D_i$ is a strict normal crossings divisor and all components $D_i$ are $k$-rational. There is linear combination $D=\sum_i m_iD_i$ with all $m_i\ge 1$ such that ${\mathcal O}_{\tilde{X}}(-D)$ is relatively ample.
We define $\tilde{X}_n$ to be scheme $\cup_i D_i$ with structure sheaf ${\mathcal O}_{\tilde{X}}/{\mathcal O}_{\tilde{X}}(-(n+1)D)$, so $\tilde{X}_0=D$, and we also define $D_{{\rm red}}$ with structure sheaf ${\mathcal O}_{\tilde{X}}/{\mathcal O}_{\tilde{X}}(-\sum_i D_i)$.
Then the functors ${\mathcal P} ic(\tilde{X}_n/k)$ and ${\mathcal P} ic(D_{ {\rm red}}/k)$, taken as a Zariski, an \'etale or a fppf functor, are representable by locally algebraic $k$-group-schemes ${\rm Pic}(\tilde{X}_n/k)$ and ${\rm Pic}(D_{{\rm red}}/k)$, so ${\rm Pic}(\tilde{X}_n)={\rm Pic}(\tilde{X}_n/k)(k), \ {\rm Pic}(D_{ {\rm red}})={\rm Pic}(D_{ {\rm red}}/k)(k)$ (see \cite[p.~273]{Kl}, \cite[Theorem~1.2]{Lip}). On the other hand, for all $n\ge 0$, and all $k$-algebras $R$, one has ${\rm Pic}(\tilde{X}_n\otimes_k R)=H^1(\tilde{X}_n\otimes_k R, {\mathcal O}^\times)$. As the relative dimension of $\sigma$ is $1$, this implies that the transition homomorphisms ${\rm Pic}(\tilde{X}_{n+1})\to {\rm Pic}(\tilde{X}_n)\to {\rm Pic}(\tilde{X}_0)\to {\rm Pic}(D_{ {\rm red}})$ are all surjective, and that ${\rm Ker}\big( {\rm Pic}(\tilde{X}_{n+1})\to {\rm Pic}(\tilde{X}_n)\big)=H^1\big(\tilde{X}_0, {\mathcal O}_{\tilde{X}_0}(-(n+1)D)\big)$. Since $-D$ is a relatively ample divisor on $\tilde{X}$, there is a $n_0\ge 0$ such that the transition homomorphisms ${\rm Pic}(\tilde{X}_n) \to {\rm Pic}( \tilde{X}_{n_0})$ are all constant for $n\ge n_0$. Since the $1$-component ${\rm Pic}^0(D_{{\rm red}})$
of ${\rm Pic}(D_{{\rm red}})$ is a semi-abelian variety, so in particular smooth, and the fibers ${\rm Pic}(\tilde{X}_n) \to {\rm Pic}(D_{{\rm red}})$ are affine \cite[p.~9,Corollaire]{Oo},
${\rm Pic}( \tilde{X}_{n_0})$ is smooth. One defines
\ga{3.1}{{\rm Pic}(\tilde{X})= {\rm Pic}( \tilde{X}_{n_0}).}
It is thus a locally algebraic smooth $k$-group-scheme.
It is an extension of $\oplus_i {\mathbb Z}[D_i]$ by its $1$-component.
Its $1$-component ${\rm Pic}^0(\tilde{X})\subset {\rm Pic}(\tilde{X}) $ is an extension of a semiabelian variety by smooth, connected commutative unipotent algebraic group over $k$.
Let $\langle D\rangle\subset {\rm Pic}(\tilde{X})$ be the subgroup-scheme spanned by those divisors with support in $D$. (In fact, $\langle D\rangle$ injects into ${\rm Pic}(D_{ {\rm red}})$ via the surjection ${\rm Pic}(\tilde{X})\to {\rm Pic}(D_{ {\rm red}})$).
It is a discrete subgroup-scheme. One sets
\ga{3.2}{{\rm Pic}(U)={\rm Pic}(\tilde{X})/\langle D\rangle.}
The Zariski tangent space at $1$ is
\ga{3.3}{H^1(\tilde{X}, {\mathcal O}_{\tilde{X}})=H^1(\tilde{X}_n, {\mathcal O}_{\tilde{X}_n})= {\rm Ker}\big({\rm Pic}(\tilde{X}_n[\epsilon])\to {\rm Pic}(\tilde{X}_n) \big)} for $n\ge n_0$, where $\tilde{X}_n[\epsilon]:=\tilde{X}_n\times_k k[\epsilon]/(\epsilon^2)$. Since ${\rm Pic}(\tilde{X})$ is smooth,
\ga{3.4}{{\rm dim}_kH^1(\tilde{X}, {\mathcal O}_{\tilde{X}})={\rm dim} \ {\rm Pic}^0(\tilde{X})={\rm Pic}^0(U).}
The last equality comes from the fact that $\langle D\rangle\subset {\rm Pic}(\tilde{X})$ is a discrete \'etale subgroup.
Recall that the surface singularity $(X,x)$ is said to be {\it rational} if $H^1(\tilde{X}, {\mathcal O}_{\tilde{X}})=0$. The definition does not depend on the choice of the resolution $\sigma: \tilde{X}\to X$ of singularities of $(X,x)$.
One has
\begin{lem} \label{lem3.1}
The following conditions are equivalent.
\begin{itemize}
\item[1)]
The surface singularity $(X,x)$ is rational.
\item[2)] ${\rm Pic}^0(\tilde{X})=0$.
\item[3)] ${\rm Pic}(U)$ is finite.
\end{itemize}
\end{lem}
\begin{proof}
The equivalence of 1) and 2) is given by \eqref{3.4}. As $\langle D\rangle \subset {\rm Pic}(\tilde{X})$ is discrete, the definition \eqref{3.2} shows that 3) implies 2). Vice-versa, assume 2) holds. Then ${\rm Pic}(\tilde{X})$ is a discrete group of finite type. Let $L\in {\rm Pic}(\tilde{X})$. Since the intersection matrix $(D_i\cdot D_j)$ is negative definite (but not necessarily unimodular), there is a $m\in {\mathbb N}\setminus \{0\}$ such that $L^{\otimes m}\in \langle D\rangle \subset {\rm Pic}(\tilde{X})$. Thus any $L\in {\rm Pic}(\tilde{X})$ has finite order in ${\rm Pic}(U)$. Since ${\rm Pic}(\tilde{X})$ is of finite type, this shows 3).
\end{proof}
\section{The Theorems}
Throughout this section, we assume $k$ to be a field, perfect if of characteristic $p>0$, $A$ to be a complete normal local $k$-algebra with maximal ideal $\frak{m}$, of Krull dimension $2$ over $k$. We set $X={\rm Spec \,} A$, $U=X\setminus \{x\}$, where $x\in X(k)$ is the rational point associated to $\frak{m}$. We say $(X,x)$ is a {\it surface singularity} over $k$. We denote by $\sigma: \tilde{X}\to X$ a desingularization such that $\sigma^{-1}(x)_{{\rm red}}=\cup_i D_i$ is a strict normal crossings divisor.
We define $H^i(Z,{\mathbb Z}_\ell(1)):=\varprojlim_n H^i(Z, \mu_{\ell^n})$ for a $k$-scheme $Z$.
\begin{thm} \label{thm4.1}
Let $(X,x)$ be a surface singularity over an algebraically closed field $k$. The following conditions are equivalent
\begin{itemize}
\item[1)] $H^1(\tilde{X},{\mathbb Z}_\ell(1))=0$.
\item[2)] $H^1(\tilde{U}, {\mathbb Z}_\ell(1))=0$.
\item[3)] There is a prime number $\ell$, different from $p$ if ${\rm char}(k) =p>0$, such that $\pi^{\rm et,ab,\ell}(U,x)$ is finite.
\item[4)] For all prime numbers $\ell$, $\pi^{\rm et,ab,\ell}(U,x)$ is finite and if ${\rm char}(k)=p>0$, then $\pi^{\rm et,ab,\ell}(U,x)=0$.
\item[5)] ${\rm Pic}^0(\tilde{X})={\rm Pic}^0(U)$ is a smooth, connected commutative unipotent algebraic group-scheme over $k$.
\item[6)] $D$ is a tree of $\mathbb{P}^1$s.
\item[7)] ${\rm Pic}^0(D_{{\rm red}})=0$.
\end{itemize}
\end{thm}
\begin{proof}
We firt make general remarks.
For any surface singularity, one has the localization sequence
\ml{4.1}{ H^1(\tilde{X}, {\mathbb Z}_\ell(1))\to H^1(U, {\mathbb Z}_\ell(1))\to\\ H^2_{D_{{\rm red}}}(\tilde{X}, {\mathbb Z}_\ell(1))\to H^2(\tilde{X}, {\mathbb Z}_\ell(1)) \to H^2(U, {\mathbb Z}_\ell(1))\to H^3_{D_{{\rm red}}}(\tilde{X}, {\mathbb Z}_\ell(1)) \to H^3(\tilde{X}, {\mathbb Z}_\ell(1)).}
By purity \cite[Theorem~2.1.1]{Fu}, the restriction map
$H^1(\tilde{X}, {\mathbb Z}_\ell(1))\to H^1(U, {\mathbb Z}_\ell(1))$ is injective, and
$H^2_{D_{{\rm red}}}(\tilde{X}, {\mathbb Z}_\ell(1))=\oplus_i {\mathbb Z}_\ell\cdot [ D_i]$. By base change, $H^i(\tilde{X}, {\mathbb Z}_\ell(1))= H^i(D_{{\rm red}}, {\mathbb Z}_\ell(1))$. Thus this group is $0$ for $i\ge 3$, equal to $\oplus_i {\mathbb Z}_\ell\cdot [ D_i]$ for $i=2$, and equal to ${\rm Pic}(D_{{\rm red}})[\ell]$ for $i=1$. In fact, since $H^2(D_{{\rm red}}, {\mathbb Z}_\ell(1))$ is torsion free, one has ${\rm Pic}(D_{{\rm red}})[\ell]={\rm Pic}^0(D_{{\rm red}})[\ell]$, where $^0$ means of degree $0$ on each component $D_i$.
Furthermore, by definition, the map
$\oplus_i {\mathbb Z}_\ell\cdot [ D_i] \to \oplus_i {\mathbb Z}_\ell\cdot [ D_i]$ is defined by $[D_i]\mapsto \oplus_j {\rm deg}{\mathcal O}_{D_j}(D_i)$. Since the intersection matrix is definite, the map is injective, with finite torsion cokernel ${\mathcal T}$. (This cokernel is 0 if and only if the intersection matrix is unimodular). Again by purity,
$H^3_{D_{{\rm red}}}(\tilde{X}, {\mathbb Z}_\ell(1)) \subset \oplus_i H^1(D_i^0, {\mathbb Z}_\ell)$ where $D_i^0=D_i\setminus \cup_{j\neq i} D_i\cap D_j$. In particular,
$H^3_{D_{{\rm red}}}(\tilde{X}, {\mathbb Z}_\ell(1))$ is torsion free.
So we extract from \eqref{4.1} for any surface singularity the relations
\ga{4.2}{H^1(\tilde{X}, {\mathbb Z}_\ell(1))\to H^1(U, {\mathbb Z}_\ell(1))={\rm Pic}(D_{{\rm red}})[\ell]={\rm Pic}^0(D_{{\rm red}})[\ell]
}
and an exact sequence
\ga{4.3}{0\to {\mathcal T}\to H^2(U, {\mathbb Z}_\ell(1))\to
H^3_{D_{{\rm red}}}(\tilde{X}, {\mathbb Z}_\ell(1))\to 0}
with finite ${\mathcal T}$ and torsion free $H^3_{D_{{\rm red}}}(\tilde{X}, {\mathbb Z}_\ell(1))$.
As ${\rm Pic}^0(D_{{\rm red}})$ is a semiabelian variety, we see that \eqref{4.2} implies that 1), 2) and 7) are equivalent conditions.
From the exact sequence
\ga{4.4}{1\to {\mathcal O}^\times_{D_{{\rm red}}}\to \oplus_i {\mathcal O}^\times_{D_i} \to \oplus_{i<j}k^\times_{D_i\cap D_j}\to 1}
one has that 6) and 7) are equivalent. Furthermore, from the structure of ${\rm Pic}(\tilde{X})$ explained in section 3, one has that 5) is equivalent to 7).
We show that 2) is equivalent to 3). The condition 2) implies that $H^1(U, \mu_{\ell^n})\subset {\mathcal T}$ for all $n\ge 0$, thus there are finitely many $\mu_{\ell^n}$ torsors on $U$. This shows 2) implies 3). On the other hand, if ${\rm Pic}^0(D_{{\rm red}})$ is not trivial, then ${\rm Pic}(D_{{\rm red}})[\ell]$ contains ${\mathbb Z}_\ell$. Thus $H^1(U, {\mathbb Z}_\ell(1))$ contains ${\mathbb Z}_\ell$ as well by
\eqref{4.2}. Thus 3) implies 2).
Since obviously 4) implies 3), it remains to see that 3) implies 4). We assume 3).
For any commutative finite $k$-group-scheme $G$, with Cartier dual $G'={\rm Hom}(G, {\mathbb G}_m)$,
one has the exact sequence
\ga{4.5}{0\to H^1(X, G')\to H^1(U, G')\to {\rm Hom}(G, {\rm Pic}(U))\to 0.}
(See \cite[III,Th\'eor\`eme~4.1]{Boutot} and \cite[III,Corollaire~4.9]{Boutot} for the $0$ on the right, which we will use only on the proof of Theorem \ref{thm4.2}, as $k=\bar k$).
We apply it for $G={\mathbb Z}/p^n$ for some $n\in {\mathbb N}\setminus \{0,1\}$.
Since ${\rm Pic}(U)$ is an extension of a discrete (\'etale) group by ${\rm Pic}^0(U)$ which is a product of ${\mathbb G}_a$s by 5),
one has ${\rm Hom}(\mu_{p^n}, {\rm Pic}(U))=0$. On the other hand, $A\xrightarrow{x\mapsto (x^{p^n}-x)}A$ is surjective, as $A$ is complete. Thus $H^1(U, {\mathbb Z}/p^n)=H^1(X,{\mathbb Z}/p^n)= 0$. This shows that 3) implies 4) and finishes the proof of the theorem.
\end{proof}
\begin{thm} \label{thm4.2}
Let $(X,x)$ be a surface singularity over an algebraically closed field $k$.
\begin{itemize}
\item[1)] If $\pi^N_{{\rm loc}}(U,X, x)$ is a finite group-scheme, $(X,x)$ is a rational singularity, in particular the dualizing sheaf $\omega_U$ has finite order.
\item[2)] If in addition, the order of $\omega_U$ is prime to $p$, then
there is $\big(h: V\to U, \pi^N(U,x),y\big) \in {\mathcal C}_{{\rm loc}}(U,x)$ such that the surface singularity $(Y,y)$ of the integral closure $\tilde{h}: Y\to X$ is a rational double point.
\item[3)] If $\pi^N_{{\rm loc}}(U,X,x)=0$, then $(X,x)$ is a rational double point.
\end{itemize}
\end{thm}
\begin{proof} We show 1).
If $\pi^N_{{\rm loc}}(U,X, x)$ is a finite group-scheme, then, by Lemma \ref{lem2.5}, the condition 3) of Theorem \ref{thm4.1} is fulfilled, thus ${\rm Pic}^0(\tilde{X})={\rm Pic}^0(U)$ is a product of ${\mathbb G}_a$s. We apply \eqref{4.5} to $G={\mathbb Z}/p^n$. If ${\rm Pic}^0(U)$ is not trivial, then ${\rm Hom}({\mathbb Z}/p^n, {\rm Pic}(U))\neq 0$ for all $n\ge 0$. Thus $U$ admits nontrivial $\mu_{p^n}$-torsors for all $n\ge 1$, which do not come from $X$. This contradicts the finiteness of $\pi^N_{{\rm loc}}(U,X,x)$. Thus ${\rm Pic}^0(U)={\rm Pic}^0(\tilde{X})=0$. We apply Lemma \ref{lem3.1} to finish conclude that $(X,x)$ is a rational singularity. Again by Lemma \ref{lem3.1}, all line bundles on $U$, in particular the dualizing sheaf $\omega_U$ of $U$, is torsion. This proves 1).
We show 2).
So there is a $M\in {\mathbb N}\setminus \{0\}$ such that $\omega^M_U\cong {\mathcal O}_U$. Choosing such a trivialization yields an ${\mathcal O}_U$-algebra structure on ${\mathcal A}=\oplus_0^{M-1} \omega^i_U$ and thus a flat nontrivial $\mu_M$-torsor $h:V={\rm Spec \,}_{{\mathcal O}_U} {\mathcal A} \to U$. Since $(M,p)=1$, $h$ is \'etale, thus $(Y,y)$ is normal. In fact one has $Y={\rm Spec \,}_{{\mathcal O}_X} {\mathcal B}$ where ${\mathcal B}$ is the ${\mathcal O}_X$-algebra $j_*{\mathcal A}, j: U\subset X$. By duality theory, $h_*\omega_Y={\mathcal H} om_{{\mathcal O}_X}(h_*{\mathcal O}_Y, \omega_X)\cong_{{\mathcal O}_X} h_*{\mathcal O}_Y$. Let $y\in Y$ be the closed point of $Y$. Thus $(Y,y)$ is a Gorenstein normal surface singularity. On the other hand, since $h$ is a $\mu_M$-torsor, one has $\pi^N(V,y)\subset \pi^N(U,x)$, thus $\pi^N_{{\rm loc}}(V,Y,y)\subset \pi^N_{{\rm loc}}(U,X,x)$, and therefore is a finite $k$-group-scheme. Thus by 1) it is a rational singularity. Thus $(Y,y)$ is a Gorenstein rational singularity, thus is a rational double point (\cite{Durfee}).
Now 3) follows directly from 2) as $\omega_U$ has then order $1$.
\end{proof}
We now refer to \cite[Section~3]{Ar3} for the notation, and we go to Artin's list \cite[Section~4/5]{Ar3} to conclude using Theorem \ref{thm4.2} 3):
\begin{cor} \label{cor4.3}
If $\pi^N_{\rm loc}(U,X,x)=0$, then $X$ admits a finite morphism $f:\widehat{{\mathbb A}}^2\to X$. The morphism $f$ is the identity (i.e. $(X,x)$ is smooth) except possibly in the cases:
\begin{itemize}
\item[1)] ${\rm char}(k)=2$, $E^1_8, E_8^3$
\item[2)] ${\rm char}(k)=3$, $ E_8^1$
\end{itemize}
\end{cor}
\bibliographystyle{plain}
\renewcommand\refname{References}
|
\section{Introduction}\label{s:introduction}
Inspiraling massive black hole binaries (MBHBs) with masses in the range $\sim 10^4-10^{10}\,{\rm M_\odot}$ are
expected to be the dominant source of gravitational waves (GWs) at $\sim$ nHz -- mHz frequencies
\citep{Haehnelt94,Jaffe03,Wyithe03,Sesana04,Sesana05}.
The frequency band $\sim 10^{-5}\,\mathrm{Hz}$ -- $1 \,\mathrm{Hz}$
will be probed by the {\it Laser Interferometer Space Antenna} ({\it LISA}, \citealt{Bender98}), a
space-borne GW laser interferometer developed by ESA and NASA. The observational window
$10^{-9}\,\mathrm{Hz}$ -- $10^{-6} \,\mathrm{Hz}$, corresponding roughly to orbital periods 0.03 -- $30\,\mathrm{yr}$,
is already accessible using Pulsar Timing Arrays
(PTAs; e.g. the Parkes radio-telescope, \citealt{Manchester08}).
The complete Parkes PTA (PPTA, \citealt{Manchester08}), the European Pulsar Timing Array (EPTA, \citealt{Janssen08}), and NanoGrav
\citep{jen09} are expected to improve considerably
on the capabilities of these surveys, eventually joining their efforts in the
international PTA project (IPTA, \citealt{Hobbs09});
and the planned Square Kilometer Array (SKA, \citealt{Lazio09})
will produce a major leap in sensitivity.
Radio pulses generated by rotating neutron stars travel through the Galactic interstellar medium and are detected
by radio telescopes on Earth. The arrival times of pulses are fitted for a model including all the known and
measured systematic effects affecting the signal generation, propagation and detection \citep{ed06}.
Timing residuals between the observed
pulses and the best fit model, carry information on additional unmodelled effects, including the presence of GWs.
Indeed, GWs modify the propagation of radio signals from the pulsar
to the Earth \citep{Sazhin78,Detweiler79,Bertotti83,Hellings83,Jenet05},
and PTAs measure the direction dependent systematic variations in the arrival times of signals from
a sample of nearly stationary pulsars in the Galaxy distributed over the sky.
PTAs provide a direct observational window onto the MBH binary population,
and can contribute to address a number of open astrophysical questions, such as the shape
of the bright end of the MBH mass function, the nature of the MBH-bulge relation
at high masses, and the dynamical evolution at sub-parsec scales of the most
massive binaries in the Universe (particularly relevant to the so-called ``final parsec problem'',
\citealt{Milos03})
PTAs can detect gravitational radiation of two forms: (i) the stochastic GW background produced by the
incoherent superposition of radiation from the whole cosmic population of MBHBs and
(ii) individual sources that are sufficiently bright in GWs to outshine the background
(typically massive, $M \lower .75ex \hbox{$\sim$} \llap{\raise .27ex \hbox{$>$}} 10^{9}\,{\rm M_\odot}$, and ``cosmologically nearby'', $d_L \lower .75ex\hbox{$\sim$} \llap{\raise .27ex \hbox{$<$}} 3\,\mathrm{Gpc}$).
Both classes of signals are of great interest, and PTAs could lead to the discovery of systems
difficult to detect with other techniques
\citep[for alternatives using active galactic nuclei, see][and references therein]{HKM09}.
Popular scenarios of massive black hole (MBH) formation and evolution
\citep[e.g.][]{VHM03,Koushiappas06,Malbon07,Yoo07} predict frequent MBH mergers (up to
several hundreds per year), implying the existence of a large number of sub-parsec MBHBs.
The prospect for detecting GW signals using PTAs depends on the number and cosmological distribution
of MBHBs with orbital periods of 0.03 -- $30\,\mathrm{yr}$, or separations typically in the range $0.001-0.1\,\mathrm{pc}$.
The three main ingredients for calculating the GW background are
\begin{enumerate}
\item[(i)] the merger rate of MBHBs as a function of mass and redshift,
\item[(ii)] the relative time each binary spends at these separations during a merger episode
and
\item[(iii)] the amplitude of the GW signal produced by each individual stationary system.
\end{enumerate}
Recently \citet[SVC08 hereinafter]{SVC08} and \citet[SVV09 hereinafter]{SVV09},
carried out a detailed study of the expected signals
(stochastic and individual), focusing on the uncertainties related to (i).
They found that the background is affected by the
galaxy merger rate evolution along the cosmic history, the massive black hole mass function,
and the accretion history of the MBHB during a galaxy merger, and they predict a factor of
$\sim 10$ uncertainty for the characteristic strain amplitude in the range
$2\times 10^{-15}-2\times10^{-16}$, at $f=1/$yr, within the expected detection capabilities of the
complete PPTA and of the SKA. They pointed out that the GW signal can be separated into individually resolvable sources
and a stochastic background, and found the number of individually resolvable sources for a 1ns timing
precision level to be between 5 to 15, depending on the considered model.
In this paper, we examine for the first time how predictions relevant for PTA observations are
modified by the presence of ambient gas, affecting the inspiral rate of binaries during a merger episode,
ingredient (ii) above. A gaseous envelope is expected to surround the binary because
MBHBs are produced in galaxy mergers, which are known to trigger
inflows of large quantities of gas into the central region, as shown by hydrodynamic simulations
\citep{springel}. This gas, accreted onto the MBHs, is responsible for luminous AGN activity, and
is also expected to catalyze the coalescence of the new-formed MBH pair \citep[e.g.,][]{escala04, dotti07},
as described below.
The forming MBHB spirals inward initially as a result of dynamical friction on
dark matter, ambient stars, and gas \citep{BBR}. As the binary separation shrinks to sub-pc scales,
the supply rate of stars crossing the orbit decreases, and the interaction with
stars becomes less and less efficient to shrink the binary.
In gas rich mergers, the dense nuclear gas is expected to
cool rapidly and settle into a geometrically-thin
circumbinary accretion disk \citep[e.g.,][]{barnes,escala04}.
Torques from the tidal field of the binary clear a gap in the gas with a radius
less than twice the separation of the binary, and generates a spiral density wave
in the gaseous disk, which in turn drains angular momentum away from
the binary on a relatively short timescale within the last parsec, $ \lower .75ex\hbox{$\sim$} \llap{\raise .27ex \hbox{$<$}} 10^7\,$yr
(\citealt{escala05,an02,an05,dotti07,Hayasaki09}, see however \citealt{lodato09}).
Ultimately, at even smaller separations, corresponding to an orbital timescale
of $\sim$ years, the emission of GWs becomes the dominant mechanism driving the
binary to the final coalescence.
The main point of this paper, is to notice that the most sensitive PTA frequency band
corresponds to orbital separations near the transition between gas and
GW dominated evolution . {\it As binaries shrink more
quickly inwards in the gas-driven phase, the number of binaries emitting at each given separation
is decreased compared to the purely GW-driven case.} The subject of this work
is to explore how various gas-driven models modify the expectations on the GW signal
potentially observable by PTAs.
Recently, \citet[HKM09 hereafter]{HKM09}, examined the evolution of MBHBs in the gas--driven regime
for simple models of geometrically thin circumbinary disks \citep[see also,][]{syerclarke,an05}.
The interaction between the binary and the gaseous disk is analogous to type-II
planetary migration, and evolves through two main phases. First, the inspiral
is analogous to the disk--dominated type-II migration of planetary dynamics, where
the binary migrates inwards with a radial velocity equal to that of the gas accreting
towards the center. Later, as the mass of the gas within a few binary separations becomes
less than the reduced mass of the binary, the evolution slows down, and it is analogous
to the planet--dominated (or secondary--dominated) type-II migration. In both cases, the
radial inspiral rate is still much faster than in the purely GW driven
case at orbital separations beyond a few hundred Schwarzschild radii.
For standard Shakura-Sunyaev $\alpha$--disk models,
the viscosity is assumed to be proportional to the total
pressure, and is consequently very large in the radiation pressure dominated phase at small radii,
increasing the migration rate in the radiation pressure dominated phase.
On the other hand, for $\beta$--disk models, where the viscosity is proportional to the gas pressure only, the
increase of radiation pressure does not impact the viscous timescale, and the migration
rate is relatively slower in this regime.
Finally, we note that the binary-gas interaction is also significantly different for
non--steady models of accretion (\citealt{ivanov}, HKM09). In the typical secondary--dominated phase,
gas flows in more quickly then how the binary separation shrinks, and is repelled close to the
outer edge of the gap by the torques of the binary. This causes gas to accumulate near the gap,
and delays the merger of the binary relative to the steady-state models.
In this paper, we couple the HKM09 models for the migration of MBHBs in the presence of a
steady gaseous disk, to the population models derived in SVV09, and we compute the effects on GWs
at nHz frequencies. Here, we restrict to nearly circular inspirals for simplicity, using the
corresponding GW spectrum (ingredient--iii above). This assumption might be violated in
gas driven inspirals \citep{an05,cua09}, and is the subject of a future paper
(Sesana \& Kocsis 2010, in prep).
The paper is organized as follows. In Section 2 we introduce the theory of the GW signal from a MBHB population,
describing its characterization in terms of its {\it stochastic level} and of the statistics of {\it individually
resolvable sources}. In Section 3 we describe our MBHB population model, coupling models
of coalescing binaries derived by cosmological N-body simulations to a scheme for the dynamical evolution of the
binaries in massive circumbinary disks. We present in detail our results in Section 4, and in Section 5 we briefly
summarize our main findings. Throughout
the paper we use geometric units with $G=c=1$.
\section{Description of the gravitational wave signal}
The theory of the GW signal produced by the superposition of radiation from a large number of individual
sources, was extensively presented in SVC08 and SVV09, here we review the basic concepts, deriving
the GW signal for gas driven mergers.
The dimensionless characteristic amplitude of the GW background produced by a
population of binaries with a range of masses $m_1$ and $m_2$ and redshifts $z$ is given by \citep{Phinney01}
\begin{equation}
h_c^2(f) =\frac{4}{\pi f^2}
\int \mathrm{d} z\mathrm{d} m_1\mathrm{d} m_2
\, \frac{\partial^3 n}{\partial z \partial m_1 \partial m_2}
{1\over{1+z}}~{{\mathrm{d} E_{\rm gw}} \over {\mathrm{d} \ln{f_r}}}\,.
\label{e:Phinney}
\end{equation}
where $dE_{\rm gw}/d\ln{f_r}$ is the total emitted GW energy per logarithmic
frequency interval in the comoving binary rest-frame,
$f_r = (1 + z) f$ is the rest-frame frequency,
$f$ is the observed frequency, $n$ is the comoving number density
of sources, and the $1/(1+z)$ factor accounts for the redshift of the
observed GW energy. The characteristic GW amplitude $h_c$ is related to the
present day total energy in GWs as $\rho_{\rm GW}=\frac{\pi}{4}\int f h_c^2(f) \mathrm{d} f$.
\subsection{GW-driven inspirals}
To the leading quadrupole order, for circular purely GW-driven binaries orbiting far outside the innermost
stable circular orbit (ISCO), Eq.~(\ref{e:Phinney})
can be evaluated assuming that $E_{\rm gw}= E_{\rm pot} = -m_1m_2/a$ is equal to the Newtonian
potential energy of the binary, and that $f_r$ is equal to twice the Keplerian orbital frequency
\citep{Phinney01}. In this case
\begin{equation}
\frac{\mathrm{d} E_{\rm gw}}{\mathrm{d} \ln f_r} = \frac{1}{3} \mu (\pi M f_r)^{2/3} = \frac{1}{3}(\pi f_r)^{2/3} {\cal M}^{5/3}
\label{e:Egw-Phinney}
\end{equation}
for $f_r<f_{\rm ISCO}\equiv 1/(6^{3/2}\pi M)$.
Here $\mu=m_1m_2/M$, $M=m_1+m_2$, and ${\cal M}^{5/3}=\mu M^{2/3}$ are the reduced, total, and
chirp masses for a binary, and $f_{\rm ISCO}$ is the GW frequency at ISCO. Substituting into Eq.~(\ref{e:Phinney}),
\begin{equation}
h_c^2(f) =\frac{4 f^{-4/3}}{3\pi^{1/3}}
\int \mathrm{d} z\mathrm{d} m_1\mathrm{d} m_2
\, \frac{\partial^3 n}{\partial z \partial m_1 \partial m_2}
\frac{\mu M^{2/3}}{(1+z)^{1/3}}.
\label{e:hc-GWdriven}
\end{equation}
It is also useful to examine the number of MBHBs and their respective contributions
to the total signal (\citealt{Phinney01}; SVC08). Eq.~(\ref{e:hc-GWdriven}) can be rewritten as
\begin{equation}
h_c^2(f) =
\int \mathrm{d} z\mathrm{d} m_1\mathrm{d} m_2
\, \frac{\partial^4 N}{\partial m_1 \partial m_2\partial z \partial \ln f_r}
h^2({\cal M}, z, f_r),\label{e:hc_GWdrivenN}
\end{equation}
where $N$ is the number of sources, which we can calculate from a comoving merger rate density as explained below, and
\begin{equation}
h({\cal M}, z, f_r) = \frac{8}{10^{1/2}} \frac{{\cal M}}{d_L(z)} (\pi {\cal M} f_r)^{2/3},
\label{e:h(f_r)}
\end{equation}
is the sky and polarization averaged characteristic GW strain amplitude of a single binary with chirp mass
${\cal M}$, at the particular orbital radius corresponding to $f_r$.
We generate the distribution $\partial^4 N/(\partial m_1 \partial m_2\partial z \partial \ln f_r)$ corresponding to a
comoving merger rate density\footnote{In practice $\partial^4 N/(\partial m_1 \partial m_2 \partial t_r \partial V_c)$
is a function of $m_1$, $m_2$, and $z$.} $\partial^4 N/(\partial m_1 \partial m_2 \partial t_r \partial V_c)$ (see
Sec.~\ref{s:MergerRate}), assuming that the number of binaries emitting in the interval $\ln f_r$ is proportional
to the time the binary spends at that frequency,
\begin{equation}
\frac{\partial^4 N}{\partial m_1 \partial m_2\partial z \partial \ln f_r} =
\frac{\partial^4 N}{\partial m_1 \partial m_2 \partial t_r \partial V_c} \frac{\mathrm{d} V_c}{\mathrm{d} z}
\frac{\mathrm{d} z}{\mathrm{d} t_r} \frac{\mathrm{d} t_r}{\mathrm{d} \ln f_r}
\end{equation}
where $\mathrm{d} V_c/\mathrm{d} z$ and $\mathrm{d} z/\mathrm{d} t_r$ are given by the standard cosmological relations between comoving volume,
redshift, and time, given in e.g. \cite{Phinney01}. The last factor can be expressed using the residence time
$t_{\rm res} = a (\mathrm{d} a/\mathrm{d} t_r)^{-1}$ the binary spends at a particular semimajor axis as
\begin{equation}
\left|\frac{\mathrm{d} t_r}{\mathrm{d} \ln f_r}\right|
= \left|\frac{\mathrm{d} t_r}{\mathrm{d} \ln a}\frac{\mathrm{d} \ln a}{\mathrm{d} \ln f_r}\right|
= \frac{2}{3}t_{\rm res},
\label{e:dtdlnf}
\end{equation}
where Kepler's law, $a = M (\pi M f_r)^{-2/3}$, was used to obtain $\mathrm{d} \ln a/\mathrm{d} \ln f_r = 2/3$, and the
residence time for a purely GW-driven evolution is
\begin{equation}
t_{\rm res} = t_{\rm res}^{{\rm gw}} \equiv \frac{\mathrm{d} t_r}{\mathrm{d} \ln a} = \frac{5 }{64} {\cal M} (\pi {\cal M} f_r)^{-8/3}\label{e:tresgw}.
\end{equation}
In summary, the distribution of sources in Eq.~(\ref{e:hc_GWdrivenN}) becomes
\begin{equation}
\frac{\partial^4 N}{\partial m_1 \partial m_2 \partial z \partial \ln f_r}
=
\frac{2}{3}\frac{\partial^4 N}{\partial m_1 \partial m_2 \partial t_r \partial V_c} \frac{\mathrm{d} V_c}{\mathrm{d} z}
\frac{\mathrm{d} z}{\mathrm{d} t_r} t_{\rm res}.\label{e:hc_GWdrivenN2}
\end{equation}
Equations (\ref{e:hc-GWdriven}) and (\ref{e:hc_GWdrivenN}--\ref{e:hc_GWdrivenN2}) are equivalent, but
(\ref{e:hc_GWdrivenN}--\ref{e:hc_GWdrivenN2}) are practical to generate discrete Monte Carlo realizations of
a given source population. Moreover, Equations~(\ref{e:hc_GWdrivenN}--\ref{e:hc_GWdrivenN2}) provide a
transparent interpretation of (\ref{e:hc-GWdriven}). The total RMS background, $h_c \propto \sqrt{N} h \propto f^{-2/3}$,
comes about because the mean number of binaries per frequency bin is $N\propto t_{\rm res}^{{\rm gw}} \propto f_r^{-8/3}$
and each binary generates an RMS strain $h\propto f_r^{2/3}$. The scaling $h_c\propto f^{-2/3}$ is a consequence
of averaging over the local inspiral episodes of merging binaries in the GW driven regime, but is completely
independent on the overall cosmological merger history or on the involved MBH masses. The latter affects only
the overall constant of proportionality. Eq.~(\ref{e:hc-GWdriven}) also shows that this scaling constant is
insensitive to the cosmological redshift distribution of mergers, $h_c\propto(1+z)^{1/6}$, as well as the
number of minor mergers with $\mu\ll M$, once the total mass in satellites merging with BHs of mass $M$ is
fixed \citep{Phinney01}. However, the background is sensitive to the assembly scenarios of major mergers (SVV09).
More importantly, as shown in SVC08, the actual GW signal in any single realization of inspiralling binaries is
qualitatively different from $h_c\propto f^{-2/3}$, as a small discrete number of individual
massive binaries dominate the nHz GW-background, creating a very spiky GW spectrum. We further discuss the
discrete nature of the signal in Section~\ref{s:statistics} below.
\subsection{Gas-driven inspirals}
\label{s:gas}
Let us now derive the GW background for an arbitrary model of binary evolution. We can derive the background
using the residence time $t_{\rm res}$ the binary spends at each semimajor axis $a$, where $t_{\rm res} < t_{\rm res}^{{\rm gw}}$ in
the gas driven phase (cf. Eq.~[\ref{e:tresgw}] for the purely GW driven case). We define $t_{\rm res}$ for various
accretion-disk models in Section~\ref{s:gasdetails}. Generally, the emitted GW spectrum is
\begin{equation}
\frac{\mathrm{d} E_{\rm gw}}{\mathrm{d} \ln f_r} = \frac{\mathrm{d} E_{\rm gw}}{\mathrm{d} t_r} \frac{\mathrm{d} t_r}{\mathrm{d} \ln f_r} =
\frac{2}{3} \frac{\mathrm{d} E_{\rm gw}}{\mathrm{d} t_r}t_{\rm res}.
\end{equation}
The second equality follows the definition of $t_{\rm res}$ and Kepler's law (see Eq.~[\ref{e:dtdlnf}]). The emitted
power $\mathrm{d} E_{\rm gw}/\mathrm{d} t_r$ depends only on the masses and the geometry of the orbit, but is independent of the
global migration rate of the binary, and therefore it is the same as in the pure GW driven case. The effects of
migration is fully encoded in $t_{\rm res}$. Plugging back in Eq.~(\ref{e:Phinney}), the mean square signal in the gas
driven phase is
\begin{eqnarray}
h_c^2(f) =&&\frac{4 f^{-4/3}}{3\pi^{1/3}}
\int \mathrm{d} z\mathrm{d} m_1\mathrm{d} m_2
\, \frac{\partial^3 n}{\partial z \partial m_1 \partial m_2}
\frac{\mu M^{2/3}}{(1+z)^{1/3}} \nonumber\\
&&\times \frac{t_{\rm res}(M, \mu, f_r)}{t_{\rm res}^{{\rm gw}}({\cal M}, f_r)}
\label{e:hc-gasdriven}
\end{eqnarray}
A comparison with (\ref{e:hc-GWdriven}) shows that the RMS signal is attenuated in the gas driven phase by
$\sqrt{t_{\rm res}(M, \mu, f_r)/t_{\rm res}^{{\rm gw}}({\cal M}, f_r)}$. This factor is a complicated function of $f_r$, that behaves
differently for different masses and mass ratios of the binaries; the overall spectrum is no longer a powerlaw.
We generate Monte Carlo realizations of the GW signal by sampling the population of the inspiralling systems.
To do this, it is sufficient to recognize that the derivation given by Eqs.~(\ref{e:hc_GWdrivenN}--\ref{e:hc_GWdrivenN2})
remains valid in the gas drive phase if using the appropriate $t_{\rm res}(M, \mu, f_r)$, since the contribution given
by each individual source to the signal is the same. {\it The net GW spectrum changes because of a reduction in
the number of sources in the gas driven case}.
Note that the individual GW signal given by Eq.~(\ref{e:h(f_r)}) depends on three parameters only:
${\cal M}$, $z$, and $f_r$. This implies that signals from sources with the same observed frequency, redshift, and chirp
mass, $(f_r, z, {\cal M})$, but different mass ratios, $q$, are totally indistinguishable\footnote{This is true only
in the angular averaged approximation. We neglect the directional sensitivity of PTAs.}. We can make use of this
property and reduce the number of independent parameters in the distribution by integrating over the mass ratio
in Eq.~(\ref{e:hc_GWdrivenN})
\begin{equation}
\frac{\partial^3 N}{\partial {\cal M} \partial z \partial \ln f_r}
=
\int_0^1 dq
\frac{\partial^4 N}{\partial m_1 \partial m_2 \partial z \partial \ln f_r}\left|\frac{\partial(m_1,m_2)}{\partial({\cal M},q)}\right|,
\label{e:hc_GWdrivenN3}
\end{equation}
Note, that this step is different for the GW and gas driven cases, because
$\partial^4 N/(\partial m_1 \partial m_2 \partial z \partial \ln f_r)$ is proportional to $t_{\rm res}$ in
Eq.~(\ref{e:hc_GWdrivenN2}). Here $|\partial(m_1,m_2)/\partial({\cal M},q)|$ is the determinant of the
Jacobian matrix corresponding to the variable change from $(m_1, m_2)$ to $({\cal M}, q)$. With Eq.~(\ref{e:hc_GWdrivenN2}) and
(\ref{e:hc_GWdrivenN3}) we derive $\partial^3 N/(\partial {\cal M} \partial z \partial \ln f_r)$ for
any gas driven model given by $t_{\rm res}(M,\mu,f_r)$, and draw Monte Carlo
samples of inspiralling binaries from this distribution when generating the GW signal.
\subsection{Statistical characterization of the signal}
\label{s:statistics}
In observations with PTAs, radio-pulsars are monitored weekly for total periods of
several years.
Assuming a repeated observation in uniform $\Delta t$ time intervals for a total time $T$,
the maximum and minimum resolvable frequencies are $f_{\max}=1/(2\Delta t)$, corresponding to
the Nyquist frequency, and $f_{\min}= 1/T$. The observed GW spectrum is therefore discretely sampled
in bins of $\Delta f = f_{\min}$. For circular orbits, the frequency of the GWs is twice the orbital
frequency.
Let us examine whether the sources' GW frequency evolves during the observation relative to the size of the
frequency bins. Writing the frequency shift during an observation
time $T$ as $\Delta f_{\rm evol} \approx \dot{f}T = (\mathrm{d} \ln f/\mathrm{d} \ln a) (\mathrm{d} \ln a/ \mathrm{d} t) f T = \frac{3}{2} f T/ [(1+z)t_{\rm res}]$ and considering the frequency resolution bin to be $\Delta f_{\rm bin}=1/T$,
then the frequency evolution relative to the frequency resolution bin is
\begin{equation}\label{e:fevolution}
\frac{\Delta f_{\rm evol}}{\Delta f_{\rm bin}}\approx \frac{3}{2} \frac{ f T^2}{(1+z) t_{\rm res}}
= \frac{0.015}{1+z}{{\cal M}}_{8.5}^{5/3} f_{50}^{11/3}T_{10}^2 \frac{t_{\rm res}^{{\rm gw}}}{t_{\rm res}},
\end{equation}
where ${{\cal M}}_{8.5}$ is the chirp mass in units of $10^{8.5}M_{\odot}$, $f_{50}$ is the frequency in units of 50 nHz,
$T_{10}$ is the observation time in unit of 10 years, and for the second equality we have used equation~(\ref{e:tresgw}).
Equation~(\ref{e:fevolution}) shows that
typical binaries contribute to a single frequency bin as stationary sources in the GW-driven regime.\footnote{The detection of
the frequency shift of an ${{\cal M}}_{8.5}=f_{50}=z=1$ source would require an extended observation with $T \lower .75ex \hbox{$\sim$} \llap{\raise .27ex \hbox{$>$}} 35$\,yr.}
We shall demonstrate that this is also true in the gas driven case (see Figure~\ref{figtres} below).
We generate $N_k$ different realizations of the signal (usually $N_k=1000$), i.e. $N_k$ realizations of the MBHB
population, consistent with Eq.~(\ref{e:hc_GWdrivenN2}) (see Section~\ref{s:MergerRate}). Each of those consists
of $N_b\sim 10^3-10^4$ binaries producing a relevant contribution to the signal, which we label by
$({{\cal M}}_i, z_i, {f_r}_i)$, $i=1,2,\dots,N_{\rm b}$. The total signal (Eq.~[\ref{e:hc_GWdrivenN}]) in each frequency
resolution bin $\Delta f$ is evaluated as the sum of the contributions of each individual source
(see \cite{pau09} for the detailed numerical procedure)
\begin{equation}
h_c^2(f) =
\sum_i h_{c,i}^2({{\cal M}}_i, z_i, {f_r}_i),\label{e:hc_Nsum},
\end{equation}
where for each $f$ the sum is over the inspiralling sources emitting in the corresponding blue--shifted (i.e. restframe)
$\Delta f_r$ frequency resolution bin, and $h_{c,i}=h_i\sqrt{f_iT}$ is the angle and polarization averaged GW strain given
by Eq.~(\ref{e:h(f_r)}), multiplied by the square root of the number of cycles completed in the observation time. For comparison, we also
evaluate the continuous integral Eq.~(\ref{e:hc_GWdrivenN}), which represents the RMS average of Eq.~(\ref{e:hc_Nsum})
over different realizations of MBHB populations in the limit $N_k\rightarrow\infty$.
Since the mass function of merging binaries is in general quite steep,
the relative contributions of the few most massive binaries turn out to dominate the
background in each frequency bin. The total GW signal depends very sensitively on
these rare binaries, and the inferred spectrum is very spiky.
It is useful to separate the total signal $h_c$ into
a part generated by a population of GW--bright {\it individually resolvable sources},
and a {\it stochastic level} $h_{s}$, which
includes the contribution of all the unresolvable, dimmer sources. More precisely, in each frequency resolution bin,
we find the MBHB with the largest $h^2_{c,i}({{\cal M}}_i, z_i, {f_r}_i)$, and we define it {\it individually resolvable}
if its signal is stronger than the total contribution of all the other sources in that particular frequency bin. The
{\it stochastic level} is consequently defined by adding up only the unresolvable sources in Eq.~(\ref{e:hc_Nsum}).
Since the signal is dominated by few individual sources in each frequency bin, the $h_s(f)$ distribution obtained over
the $N_k$ realizations is far from being Gaussian or just even symmetric. To give an idea of the uncertainty range of
$h_s(f)$, we calculate the $10\%$, $50\%$ and the $90\%$ percentile levels of the $h_s(f)$ distribution of the $N_k$
different realizations.
The separation of individually resolvable sources is useful for several reasons (see SVC; SVV09).
First, it is useful from a statistical point of view for understanding the variance of
the expected GW spectrum among various realizations of the inspiralling MBHBs.
The discrete nature of the resolvable sources allows a different statistical analysis than for
the smooth background level corresponding to the stochastic level, $h_s(f)$.
The individually resolvable signals could also be important observationally.
A sufficiently GW--bright resolvable binary allows to measure the GW polarization using PTAs,
and give information on the sky position of the binary \citep{SV10},
which might be used to search for direct electromagnetic signatures
like periodically variable AGN activity (HKM09).
A coincident detection of GWs and electromagnetic emission of the same
binary would have far reaching consequences in fundamental physics, cosmology, and black hole physics
\citep{Kocsis08}.
\subsection{Timing Residuals}
\label{s:timing}
In general, the characteristic GW amplitudes (either of a stochastic background or of a resolvable source) can be
translated into the pulsar timing language by converting $h_c(f)$ into a ``characteristic timing residual'' $\delta t_c(f)$
corresponding to the sky position and polarization averaged delay in the time of arrivals of consequent pulses due to GWs,
\begin{equation}
\delta t_c(f)=\frac{h_c(f)}{2\pi f}.
\label{e:t_c}
\end{equation}
The pulsar timing residuals expected from an individual stationary GW source is derived in section 3
of SVV09 in detail. The corresponding measurement can be represented, in the time domain, with a residual:
\begin{equation}
\delta t(t) = r(t) + \delta t_{\rm N}(t)
\end{equation}
where $r(t)$ is the contribution due to the GW source (which accumulates continuously with observing time $t$, see below),
and $\delta t_{\rm N}(t)$ represents random fluctuations due to noise.
The latter is the superposition of the intrinsic noise in the measurements and the GW stochastic level
from the whole population of MBHBs, with a root-mean-square (RMS) value
\begin{equation}
\delta t_\mathrm{N,rms}^2(f) = \langle \delta t_\mathrm{N}^2(f) \rangle = \delta t_\mathrm{p}^2(f) + \delta t_\mathrm{s}^2(f).
\label{e:noise}
\end{equation}
where $t_\mathrm{p}(f)$ is the RMS instrumental and astrophysical noise corresponding to
the given pulsar, and $\delta t_\mathrm{s}(f)=h_s(f)/(2\pi f)$ is due to the RMS stochastic GW background of unresolved MBHBs,
as defined in the previous section.
The sky angle and orbital orientation-averaged signal-to-noise ratio (SNR) at which one MBHB, radiating at (GW)
frequency $f$, can be detected using a {\it single} pulsar with matched filtering is
\begin{equation}
{\rm SNR}^2 = \frac{\delta t_\mathrm{gw}^2(f)}{\delta t_\mathrm{N,rms}^2(f)}\,.
\label{e:snr}
\end{equation}
Here $\delta t_\mathrm{gw}(f)$ is the root-mean-squared timing residual signal resulting from GWs emitted by the
individual stationary source over the observation time $T$ defined as:
\begin{equation}
\delta t_\mathrm{gw}(f) =\sqrt{\frac{8}{15}}\frac{h({\cal M}, z, f)}{2\pi f} \sqrt{f T}
\label{e:deltatgw}
\end{equation}
where $h({\cal M}, z, f)$ is the angle and polarization averaged GW strain amplitude given by Eq.~(\ref{e:h(f_r)}), the
prefactor $\sqrt{8/15}$ averages the observed signal over the ``antenna beam pattern'' of the array (Eq. (21)
in SVV09\footnote{Note that that the square root in the prefactor $\sqrt{8/15}$ is missing in Eq. (20) of SVV09
because of a typo there.}), and the $\sqrt{fT}$ term accounts for the residual build-up with the number of cycles.
For $N_{\rm p}$ number of pulsars, the total detection SNR, of an individually resolvable MBHB is the RMS of the
contributions of individual pulsars given by (\ref{e:snr}). For $N_{\rm p}$ identical pulsars, the effective
noise level is therefore attenuated by $N_{\rm p}^{-1/2}$.
In the following we will represent the overall GW signal and the stochastic background by using either their
characteristic amplitudes, $h_c(f)$ and $h_s(f)$, or the corresponding characteristic timing residuals,
$\delta t_c(f)$ and $\delta t_s(f)$, according to equation (\ref{e:t_c}). We study the detection significance of individually
resolvable sources and the distribution of their numbers as a function of the induced $\delta t_{\rm gw}$. For each Monte Carlo
realization of the emitting MBHB population, we count the cumulative number of all ($N_t$) and resolvable ($N_r$) sources
above $\delta t_{{\rm gw}}$ as a function of $\delta t_{\rm gw}$:
\begin{equation}
N_{t/r}(\delta t_\mathrm{gw}) = \int_{\delta t_\mathrm{gw}}^\infty \frac{\partial N_{t/r}}{\partial \delta t_\mathrm{gw}'}\delta t_\mathrm{gw}'\,,
\label{e:N}
\end{equation}
where the integral is either over all sources or only the individually resolvable sources (i.e. restricted to those
that produce residuals above the RMS stochastic level, see Sec.~\ref{s:statistics}).
\section{The emitting binary population}
\label{s:Population}
We calculate the GW signal by generating a catalogue of binaries consistent with
Eq.~(\ref{e:hc_GWdrivenN2}). This requires (i) a model for the
comoving merger rate density of coalescing MBHBs,
$\partial^4 N/(\partial m_1 \partial m_2 \partial t_r \partial V_c)$, and
(ii) a model for the evolution of individual inspiralling binaries $t_{\rm res}(M, \mu, f_r)$.
These two items are the subjects of the next two subsections below.
\subsection{Population of coalescing massive black hole binaries}
\label{s:MergerRate}
We use the population models described in section 2 of SVV09,
the reader is deferred to that paper for full
details. We extract catalogs of merging binaries from the semi-analytical
model of \cite{Bertone07} applied to the Millennium run \citep{springel}.
We then associate a central
MBH to each merging galaxy in our catalogue.
We explored a total of 12 models, combining four $M_{\rm BH}$-bulge
prescriptions found in the literature with three different accretion scenarios
during mergers. The twelve models are listed in table 1 of SVV09.
In the present study we shall use the Tu-SA population model as our default case.
In this model, the MBH masses in the merging galaxies correlate with the
masses of the bulges following the relation reported in \cite{Tundo07},
and accretion is efficient onto the more massive black hole, {\it before}
the final coalescence of the binary.
Since the comparison among different MBHB population models is not the main purpose
of this study, we will present results only for this model. However, we tested our
dynamic scenario on other population models presented in SVV09,
finding no major differences for alternative models.
Assigning a MBH to each galaxy, we obtain a catalogue of mergers
labelled by MBH masses and redshift. From this, we generate the merger rate per comoving volume,
$\partial^4 N/(\partial m_1 \partial m_2 \partial t_r \partial V_c)$. In practice, due to the
large number of mergers in the simulation, this is a finely resolved continuous function,
describing the merger rate density as a function of $m_1$, $m_2$, and $z$.
After plugging into Eqs.~(\ref{e:hc_GWdrivenN2})
and (\ref{e:hc_GWdrivenN3}), we can obtain the continuous distribution
$\partial^3 N/(\partial {\cal M} \partial z \partial \ln f_r)$ one would observe in an ``ideal snapshot'' of the
whole sky. We then sample this distribution to generate random Monte Carlo realizations
of the GW signal. In summary, the
chosen MBHB population model fixes the cosmological merger history, i.e. the function
$\partial^4 N/(\partial m_1 \partial m_2 \partial t_r \partial V_c)$, and ``Monte Carlo
sampling'' refers to first choosing a realization of the cosmological merger history
(i.e. generate $N_b$ number of binary masses and redshifts\footnote{Here $N_b$ is chosen randomly
for each distribution, it can vary for different realizations of the population according to a Gaussian
distribution with $\sigma=1/\sqrt{\langle N_b\rangle}$ around the mean $\langle N_b\rangle$.})
and then assigning an orbital frequency (or time to merger) to each binary.
In addition to our fiducial merger rate, we also examine for the first time the situation where minor mergers
do not contribute to the coalescence rate of the central MBHs. This is motivated by recent
numerical simulations indicating that minor mergers with mass ratios $q<0.1$ lead to tidal
stripping of the merging satellite, and the resulting core does not sink efficiently to the
centre of the host galaxy \citep{cal09}. We identify the mass ratio of merging galaxies using the mass
of the stellar components, to avoid complications due to the tidal stripping of merging dark matter
halos.
In practice, we find that suppressing all the minor mergers does not affect the resulting GW signal,
implying that the contribution of mergers involving dwarf galaxies is negligible.
\subsection{Binary evolution in massive circumbinary disks}
\label{s:gasdetails}
We adopt the simple analytical models of HKM09 to describe the dynamical evolution
of MBHBs in a geometrically thin circumbinary accretion disk. {Here we provide
only a short summary of the gas-driven models, highlight their main assumptions,
and defer the reader to section 2 of HKM09 and references therein for more details.
For the typical MBH masses
and separations we are considering, the tidal torque from the binary dominates
over the viscous torque in the disk, opening a gap in the gas distribution. A spiral density
wave is excited in the disk which torques the binary and pushes it inward. First, when the
binary separation is relatively large, and the local disk mass is larger than the mass
of the secondary, then the
secondary migrates inward with the radial velocity of the accreting gas,
\begin{equation}\label{e:tnu}
t_{\rm res}^{\nu} = -\frac{a}{\dot{a}_{\nu}} = \frac{2\pi r_0^2 \Sigma_0}{\dot M}
\end{equation}
where $r_0$ is the radius of the gap (typically $r_0\sim 1.5 a$ where $a$ is the semimajor
axis of the binary), $\Sigma_0$ is the local surface density of the disk with no secondary,
and $\dot M$ is
the accretion rate far from the binary (the $\nu$ index denotes viscous evolution).
This is analogous to disk dominated Type-II planetary migration.
More typically, however, the binary is more massive than the local
disk mass. In this case, analogous to secondary-dominated Type-II migration, the
angular momentum of the binary is absorbed less efficiently by the gas outside the gap,
and the evolution slows down according to \citep{syerclarke}
\begin{equation}\label{e:tSC}
t_{\rm res}^{\rm SC} = -\frac{a}{\dot{a}_{\rm SC}} = q_B^{-k_1} t_{\nu}
\end{equation}
where $q_B$ is a measure of the lack of local disk mass dominance
\begin{equation}
q_B = \frac{4\pi r_0^2 \Sigma_0}{\mu} = \frac{2 \dot M}{\mu} t_\nu,
\end{equation}
which is less than unity in this case;
$\mu$ is the reduced mass of the binary,
and $k_1$ is a constant that depends on the surface density -- accretion rate
powerlaw index. Here $k_1=0$ if $q_B>1$; otherwise,
$k_1=7/17$ at large separations, if the disk opacity at the gap boundary is dominated by free-free absorption,
and $k_1=3/8$ closer in, if the opacity is dominated by electron scattering.
{ In the secondary-dominated regime, eq.~(\ref{e:tSC}) assumes that
the migration is very slow, so that the binary can influence the surface density very far
upstream, and an approximate steady state is reached where the gas density
is enhanced outside the gap by a factor $q_B^{-k}$
relative to the gas density with no secondary \citep{syerclarke}. To examine the
sensitivity of our conclusions to these assumptions, we consider an additional
nonsteady model of \citet{ivanov}. In their model, which is also axisymmetric by
construction, they assumed that the binary torques are
concentrated in a narrow ring near the edge of the cavity, which results in a
time-dependent pile up of material near the cavity.
The migration slows down further relative to $t_{\rm res}^{\rm SC}$ as
\begin{equation}\label{e:tIPP}
t_{\rm res}^{\rm IPP} = \frac{\mu}{2 \dot M} \left(\frac{a}{a_0}\right)^{1/2}
\left\{\frac{1}{1 + \delta \left(1 - \sqrt{a/a_0}\right)}\right\}^{k_2},
\end{equation}
where $\dot M$ is the constant accretion rate far outside the binary,
$a_0$ is the initial semimajor axis where the local disk mass is just
equal to the secondary (i.e. $q_B=1$), $\delta$ and $k_2$ are
constants which depend on the density--accretion rate profile of the disk
($\delta =4.1$ and $6.1$, $k_2=0.29$ and $0.26$ for free-free and electron scattering opacity,
respectively). { Note that as the separation decreases well below its initial value, $a\ll a_0$,
the curly bracket is $a$--independent,
implying that $t_{\rm res}^{\rm IPP} \propto a^{1/2} \propto t_{\rm orb}^{1/3}$.}
Since the migration rate is proportional to the gas surface density and accretion rate, it is
sensitive to the structure of the accretion disk. Following \citet{syerclarke},
we estimate $\Sigma_0$ with that of a steady accretion disk of a single accreting
BH, but consider two different models for that, $\alpha$ and $\beta$--disks (see HKM09 for
explicit formulae). }
For the classic \cite{ss73} $\alpha$-disk,
the viscosity is proportional to the total (gas+radiation) pressure of the
disk.
Until very recently, this model, if radiation pressure dominated,
has been thought to be thermally and viscously unstable \citep{le74,piran78}.
In the alternative $\beta$-model, the viscosity
is proportional to the gas pressure only\footnote{The name comes from the
definition of viscosity $\nu\propto \alpha p_{\rm gas}^{\beta}p_{\rm tot}^{1-\beta}$
where $\beta=1$ for the $\beta$--model, while $\beta=0$ for the
$\alpha$ model. In both cases $\alpha$ is a free model parameter.}, and
it is stable in both sense. The nature of viscosity is not well understood to predict
which of these prescriptions lies closer to reality.
Recent numerical magneto-hydrodynamic simulations \citep{hirose09}
suggests that the thermal instability is avoided in radiation pressure
dominated situations because stress fluctuations lead
the associated pressure fluctuations, and seem to favor the $\alpha$
prescription over the $\beta$--model.
We carry out all calculations for both models, but consider the $\alpha$ prescription
as our fiducial disk model.
In both cases, the model is uniquely determined by three parameters:
the central BH mass, the accretion rate { far from the binary} $\dot M$, and
the $\alpha$ viscosity parameter. The exact value of these parameters
is not well known. Observations of luminous AGN
imply an accretion rate around $\dot m = \dot{M}/\dot{M}_{\rm Edd} = (0.1$--$1)$
with a statistical increase towards higher quasar luminosities \citep{Kollmeier06,Trump09}.
Here $\dot{M}_{\rm Edd}=L_{\rm Edd}/(\eta{\rm c}^2)$
is the Eddington accretion rate for $\eta=10\%$ radiative efficiency, where
$L_{\rm Edd}$ is the Eddington luminosity. Observations of outbursts in binaries
with an accreting white dwarf, neutron star, or stellar black hole imply
$\alpha=0.2$--$0.4$ \citep[and references therein]{Dubus01,King07}.
Theoretical limits based on simulations of magneto-hydrodynamic turbulence
around black holes are inconclusive, but are consistent with
$\alpha$ in the range 0.01--1 \citep{Pessah07}. It is however unclear whether
these numbers are directly applicable to circumbinary MBH systems.
The binary exerts a torque that pushes the gas away on average, and
consequently might reduce the accretion rate (see \citealt{Lubow99,LubowDAngelo06},
in the planetary context, and \citealt{mm08,cua09}, for MBH binaries).
We explore several choices covering all 6 combinations with
$\dot{m}=\{0.1, 0.3\}$ and $\alpha=\{0.01, 0.1, 0.3\}$ for both $\alpha$ and $\beta$--disks,
respectively. Motivated by the considerations outlined above,
we highlight the $\alpha$-disk with $\dot{m}=0.3$ and $\alpha=0.3$ as our default
model.
{ We regard $t_{\rm res}^{\rm SC}$ and $t_{\rm res}^{\rm IPP}$ as lower and upper limits of the
true residence time during secondary dominated gas driven migration. However,
we caution that these estimates are subject to many uncertainties related to
the complexity of binary accretion disk and migration physics.
Other models for the binary--disk interaction \citep[i.e.][]{Hayasaki09,Hayasaki10}
lead to a slower migration rate.
Two dimensional hydrodynamic simulations show that the flow into the gap is
generally non-axisymmetric, and the gap and the binary orbit becomes eccentric
\citep{LubowDAngelo06,mm08,cua09}. The accretion rate and inspiral velocity
in reality may be higher than in the axisymmetric approximation.
Resonant interactions may lead an enhanced angular momentum transport
\citep{gt79}. For unequal masses or thick disks, there may be
significant inflow into the gap. In this case, corotation torques need to be considered,
and the migration is expected to be much faster inward or outward, analogous to
Type I planetary migration \citep{ttw02}. Disk thermodynamics
may also significantly influence the migration rate \citep{Paardekooper1,Paardekooper2}. }
These steady state accretion disk models also assume that the self-gravity of the disk is
negligible, the disk is optically thick, and the temperature of the gas is large enough
($T \lower .75ex \hbox{$\sim$} \llap{\raise .27ex \hbox{$>$}} 10^4\,$K) that the opacity corresponds to that of ionized gas. These
assumptions break down at radial distances corresponding to orbital periods of
around $t_{\rm orb} \lower .75ex \hbox{$\sim$} \llap{\raise .27ex \hbox{$>$}} (5$--$10)\,\mathrm{yr}$ for BHs with masses $M \lower .75ex \hbox{$\sim$} \llap{\raise .27ex \hbox{$>$}} (10^8$--$10^9)\,{\rm M_\odot}$
(see Fig. 1 of HKM09). Thus, these disk models are self-consistent for the
relatively short period binaries, but become suspect for binaries emitting at
the low frequency edge of the PTA window.
Another shortcoming of these gas-driven migration models, is that they were derived assuming
the disk surface density is a decreasing function of radius. This affects the specific values of $k_1$ and $k_2$
in eqs.~(\ref{e:tSC}) and (\ref{e:tIPP}). This assumption is safe for $\beta$--disks in general, as well as for
$\alpha$--disks if the gas pressure dominates over radiation pressure, but it is violated for radiation
pressure dominated $\alpha$--disks.{ However,
gas-driven models with binary separation in the PTA frequency band and large masses $M>10^8\,{\rm M_\odot}$
are in fact radiation pressure dominated near the gap.\footnote{Note that the most massive binaries, however, are GW driven, and are not sensitive to the accretion disk. } }
It is unclear how $k_1$ and $k_2$ change for radiation pressure dominated $\alpha$--disks.
As an approximation, following \citet{an05} and HKM09, we extrapolate the values of the gas pressure
dominated outer regions.
\begin{figure}
\centering
\mbox{\includegraphics[width=84mm]{fig1.eps}}
\caption{Residence time, $t_{\rm res}$, as a function of the binary period. Different
panels correspond to
selected model parameters as labelled.
In each panel the thick curves represent the residence time for the
gas driven dynamics according to HKM09.
Solid, long--dashed and short--dashed curves are for mass ratios $q=1, 0.1, 0.01$ respectively. The
two thin dotted vertical lines approximately bracket the PTA observable window.
The extrapolated pure GW-driven evolution ($t_{\rm res}^{\rm gw}\propto t_{\rm orb}^{8/3}$)
is shown as thin lines, for comparison.
The ratio $t_{\rm res}/t_{\rm res}^{\rm gw}$ gives the relative
decrease in the number of binaries in the gas dominated case compared to the GW driven
case.
}
\label{figtres}
\end{figure}
{
Substituting the surface density profiles $\Sigma_0(r)$ of
$\alpha$ and $\beta$--disks in eqs.~(\ref{e:tnu}--\ref{e:tSC}),
we obtain the inspiral rate of binaries as they evolve
from large to small radii in a gaseous medium. Figure~\ref{figtres} shows the
corresponding residence times for selected MBH masses
and disk models. The binary migration history, can be divided in three phases.
\begin{enumerate}
\item Initially, migration occurs on the viscous timescale $t_{\rm res}^{\nu} \propto t_{\rm orb}^{5/6}$.
\item Then as the disk mass decreases below the secondary mass, the gas-driven migration slows
down according to $t_{\rm res}^{\rm SC}$ or $t_{\rm res}^{\rm IPP}$. The migration rate in this case is nonuniform,
it changes as the local accretion disk structure varies due to changes in the dominant source of opacity
(free-free for large separation to electron-scattering for smaller separations) and
the source of pressure (thermal gas pressure at large $a$ to radiation pressure at small $a$).
In particular for the steady \citet{syerclarke} models, $t_{\rm res}^{\rm SC}\propto t_{\rm orb}^{25/51}$--$t_{\rm orb}^{7/12}$
with free-free to electron scattering opacity for thermal pressure support, and
$t_{\rm res}^{\rm SC}\propto t_{\rm orb}^{35/24}$--$t_{\rm orb}^{7/12}$,
for radiation pressure support for $\alpha$ to $\beta$ disks, respectively;
while for the time-dependent models of \citet{ivanov}, the accretion rate approaches $t_{\rm res}^{\rm IPP} \propto t_{\rm orb}^{1/3}$.
Note that the residence time in this regime is generally smaller for $\alpha$--disks than for $\beta$--disks.
This is explained by the fact that the overall surface density is smaller for radiation pressure
dominated $\alpha$--disks, and so to achieve the same net accretion rate, the radial gas inflow
velocity is larger for $\alpha$--disks. The migration rate is a monotonic function of the radial
inflow velocity (see eqs.~\ref{e:tnu} and \ref{e:tSC}), so that the binary is pushed in more
quickly for $\alpha$--disks.
\item Finally, when the separation is sufficiently reduced, the emission of gravitational
waves becomes very efficient and determines the inspiral rate according to
$t_{\rm res}^{\rm gw}=a/\dot a_{\rm gw} \propto a^4 \propto t_{\rm orb}^{8/3}$
(see eq.~\ref{e:tresgw}).
\end{enumerate}
}
Note that { for circular orbits, the GW frequency (in the binary restframe)} is simply $f_r = 2/t_{\rm orb}$.
The number of binaries at any given $t_{\rm orb}$ is proportional to the residence time $t_{\rm res}$. Therefore,
the decrease in $t_{\rm res}$ in the gas driven regime compared to the GW driven case implies a decrease
in the population of MBHs, which ultimately leads to the attenuation of
the low frequency end of the observable total GW spectrum. The RMS GW spectrum averaged
over the whole population of binary inspiral episodes is no longer a powerlaw.
It is interesting to examine the contributions of
various evolutionary phases according to Eq.~(\ref{e:hc-gasdriven}),
$h_c\propto f^{-2/3} \sqrt{t_{\rm res}/t_{\rm res}^{\rm gw}}$. If all binaries were
in the GW driven phase then $h_c^{\rm gw}\propto f^{-2/3}$. If all were in
the secondary-dominated Type-II migration regime with a steady radiation pressure dominated disk\footnote{
Most of the gas driven binaries contributing to the background at large frequencies are in this regime.}
$h_c^{\rm SC}\propto f^{-1/16}$--$f^{3/8}$ for $\alpha$ and $\beta$ disks, respectively,
further out $h_c^{\rm SC}\propto f^{3/8}$--$f^{43/102}$ in the gas pressure dominated regime
(with electron scattering versus free-free opacity), { while $h_c^{\rm IPP}\propto f^{1/2}$ asymptotically for the
nonsteady models}, and finally
$h_c^{\nu}\propto f^{1/4}$ in the disk-dominated Type-II migration regime. In any case, the
GW spectrum $h_c(f)$ is much shallower in the gas driven phase.
Note the gas driven phase contributes a nearly flat or an
{\it increasing} spectrum $h_c(f)$, very different from the
nominal $f^{-2/3}$ GW driven case. { In general, the orbital separation for more massive objects at a given $f_r$
is smaller in terms of Schwarzschild radii. Since the transition to gas driven migration}, when expressed in
Schwarzschild radii, is roughly independent of mass, it follows then at any given frequency bin the more massive
objects are typically GW driven and lighter objects are gas driven. The total average spectrum\footnote{if averaging
over each binary episode but not over the cosmological merger tree} assuming only wet mergers\footnote{Here
``wet'' refers to gas rich mergers where the dynamics of the binary { is driven by both the circumbinary disk
and GW emission depending on the binary separation, while ``dry'' refers to the case where the binary dynamics
is driven by GW emission only at all radii.} Note that in both cases we neglect interactions with stars.},
is between $f^{-2/3}$ and $f^{1/2}$ depending on the ratio of GW driven to gas driven binaries.
Note however, that the individually resolvable
sources are typically very massive and are in the GW-driven phase for the relevant range of binary separations, and
therefore their properties should not be modified by gas effects.
\section{Results}
\subsection{Description of the signal}
\begin{figure*}
\centering
\resizebox{\hsize}{!}{\includegraphics[clip=true]{fig2a.eps}
\includegraphics[clip=true]{fig2b.eps}}
\caption{{Components} of the GW signal from a population of inspiralling MBHBs. In the left panel we
consider {all} binaries embedded in a gaseous $\alpha$-disk for our default model (all mergers wet), while
in the right panel {all} binaries are purely GW driven (all mergers dry). In each panel, the smooth
solid line is the RMS total characteristic GW strain $h_c$ using the integral expression
(\ref{e:hc_GWdrivenN}--\ref{e:hc_GWdrivenN2}) (which correspond to an average over $N_k\rightarrow\infty$
Monte Carlo realizations), the two dashed black lines represent the RMS total signal (upper) and the RMS
stochastic background level (lower) averaged over $N_k=1000$ Monte Carlo realizations respectively. The jagged
blue line displays a random Monte Carlo realization of the GW signal. The small black and red triangles show the
contributions of the brightest and resolvable sources in each frequency bin respectively. The jagged red line is the stochastic
GW background for this realization, i.e., once the resolvable sources in each frequency bin are subtracted. The
green dots label all the systems producing an RMS residual $t_{\rm gw}>0.3\,$ns over $T=10$\,years. The dotted
diagonal lines shows constant $t_{\rm gw}$ levels as a function of frequency. An observation time of 10 years is assumed.
}
\label{f1}
\end{figure*}
As stated in Sec. 2.3, the relevant frequency band for pulsar
timing observations, assuming a temporal observation baseline $T$ and a time interval between subsequent observations
$\Delta t$, is between $f_{\rm min}\approx 1/T$ and $f_{\rm max}\approx 1/(2\Delta t)$
with a resolution $\Delta{f}\approx 1/T$. In our calculations we assume a default duration of $T=10$\,yr for the
PTA campaign with $\Delta t\approx 1$ week. This gives $f_{\rm min}\approx 3\times 10^{-9}$\,Hz, $f_{\rm max}\approx 10^{-6}$Hz
and $\Delta{f}\approx 1/T \approx 3\times 10^{-9}\,$Hz.
The simulated signal is computed by doing a Monte Carlo sampling
of the distribution $\partial^3N/(\partial z\partial {\cal M} \partial \ln f_r)$, and adding the GW contribution of
each individual source. In each frequency bin, we identify the individually resolvable sources and the stochastic background.
We repeat this exercise $N_k=1000$ times for the 12 steady disk models defined in Sec.~\ref{s:gasdetails} and for the purely
GW--driven case. In addition, we also calculate the GW signal using the integral expression (\ref{e:hc_GWdrivenN}--\ref{e:hc_GWdrivenN2})
using the continuous distribution function, which corresponds to the RMS average of the GW signal in the $N_k\rightarrow\infty$ limit.
The GW signal and its most important ingredients are plotted in Figure~\ref{f1}. The left panel shows
results for gas--driven inspirals using our default $\alpha$-disk (i.e. ``gas on'' -- all mergers wet), the right panel shows
results for purely GW--driven inspirals for comparison (i.e. ``gas off'' -- all mergers dry) for the same underlying cosmological
MBHB coalescence rate. A randomly selected Monte Carlo realization of the signal is depicted as a dotted blue jagged line.
The green dots represent the contributions of individual binaries; systems producing
$\delta t_{{\rm gw}}(f)>0.3\,$ns timing residual are shown.
In each frequency bin, the brightest source is marked by a black triangle. The individually resolvable
sources are marked by superposed red triangles, and the stochastic level from all unresolvable sources is
shown with a solid red jagged line. Clearly, the GW signal for any single realization is far
from being smooth. The noisy nature of the signal is due to rare massive binaries rising well above
the stochastic level.
Solid black curves in Figure~\ref{f1} show $h_c(f)$, the RMS value of the GW signal averaged over the merger
episodes, using the integral expression (\ref{e:hc_GWdrivenN}--\ref{e:hc_GWdrivenN2}). The upper black dashed
curve is the GW signal, averaged over $N_k=1000$ Monte Carlo realizations of the merging systems. This is
still noisy, due to the finite number of realizations used, but is consistent with the integrated average
shown by the solid black curve. The lower black dashed curve marks the RMS stochastic component,
$h_{s,{\rm rms}}(f)$, averaged over the same $N_k=1000$ realizations.
Figure~\ref{f1} shows that the gaseous disk greatly reduces the number of binaries at small frequencies compared
to the GW--only case, as gas drives the binaries in quickly towards the final coalescence. Consequently, the signal is more spiky
in the presence of gas. There is a clear flattening of the RMS spectrum at low frequencies ($f<1/$yr) in the
gas-driven case compared to the purely GW-driven case. The stochastic level is more suppressed in the wet-only
case and is much steeper than $f^{-2/3}$ at high frequencies. Despite the spiky signal in a given Monte Carlo
realization, the overall shape of the background is well recognizable in the GW driven model, while the characterization
of the global shape of the signal in the gas driven case appears to be less viable.
Gas driven migration becomes more and more prominent at large binary separations, corresponding to large orbital times
or small GW frequencies. Therefore, to check the ultimate maximum impact of gas effects on future PTA detections, we
simulated the spectrum for a hypothetical $T = 100$\,yr observation baseline. Figure \ref{f100yr} shows a realization
of the spectrum for such an extended observation, assuming our default gas model (i.e. all mergers wet).
Given the extended temporal baseline, the minimum observable frequency is pushed down to $\sim3\times10^{-10}\,\mathrm{Hz}$,
where the spectrum of gas driven mergers is considerably flatter. Moreover, the
frequency resolution bin is then narrower, the number of sources per bin is much smaller, and more sources
become resolvable. At the smallest frequencies, the induced timing residual can be higher
than 1$\mu$s and more than 50 sources will be individually resolvable at 1ns precision level, some of them with an SNR as high as 100.
These numbers are not severely modified if considering purely GW--driven dry mergers only (see Fig.~\ref{f:N_resolvable(t_gw)} below).
{Interestingly, due to the high frequency resolution of such an extended observation,
the GW frequency of some of the resolvable binaries may evolve significantly during the observation
(e.g. a binary with masses $m_1=m_2=5\times 10^8\,{\rm M_\odot}$ or higher are nonstationary
relative to the bin size at frequencies above $f \lower .75ex \hbox{$\sim$} \llap{\raise .27ex \hbox{$>$}} 40\,$nHz, see equation~[\ref{e:fevolution}]). Therefore it may be possible
to detect the frequency evolution of individually resolvable binaries during such an extended monitoring campaign. }
However, this
computation is idealized: it is questionable if millisecond pulsars can maintain a ns timing stability over such
a long timescale, nonetheless, it points out the enormous capabilities of long term PTA campaigns.
\begin{figure}
\centering
\mbox{\includegraphics[width=84mm]{fig3.eps}}
\caption{Same as left panel of Figure~\ref{f1}, but for an observation lasting 100 years (Averages over the 1000
realizations are not shown in this case for clarity).}
\label{f100yr}
\end{figure}
\subsection{Individually resolvable sources}
Let us now examine the prospects for detecting individual sources using PTAs.
How many sources are expected to be individually resolvable? How significant is their detection?
As explained in Sec.~\ref{s:timing}, the
signal from a specific binary can be detected using a PTA
if the corresponding timing residual $\delta t_{{\rm gw}}^2$ is above the RMS noise level $\delta t^2_{\rm N, rms}$ characterizing the PTA
given by Eq.~(\ref{e:snr}).
We should notice, however, that this estimate of the number of individually resolvable sources
is conservative and is likely to provide only a {\it lower limit} for the following reasons.
Firstly, we average over the sky position of the binaries and pulsars, while in reality, it may be possible
to take advantage of the different GW polarization and amplitude generated by sources in different sky positions to deconvolve their
signal (even if they have similar strength and frequency). Secondly, the brightest source identification algorithm can
be implemented recursively, after an accurate subtraction of the identified sources. However, we find that, especially
at low frequencies, the distribution of GW source amplitudes for various binaries in a single frequency bin is not
strongly hierarchical, so that a recursive brightest source finding algorithm shouldn't
increase significantly the number of resolvable systems.
We present here results both in terms of the total number ($N_t$) and resolvable systems ($N_r$, see Eq.~\ref{e:N}).
\begin{figure*}
\centering
\mbox{\includegraphics[width=144mm]{fig4.eps}}
\caption{{\it Top-left panel}: characteristic amplitude of the
timing residuals $\delta t_\mathrm{gw}$ (equation (\ref{e:deltatgw})) as a function of frequency; the dots
are the residuals generated by individual sources and the solid line is the estimated {\it stochastic level}
of the GW signal. {\it Top-right panel}:
distribution of the number of total (dotted lines) and resolvable (solid lines) sources per logarithmic
redshift interval as a function of redshift, generating a
$\delta t_\mathrm{gw}>1$ns.
{\it Bottom-left panel}:
distribution of the number of total (dotted lines) and resolvable (solid lines) sources per logarithmic
frequency interval as a function of the GW frequency, generating a
$\delta t_\mathrm{gw}>1$ns.
{\it Bottom-right panel}: distribution of the total (dotted lines) and individually resolvable
(solid lines) number of sources per logarithmic chirp mass interval
as a function of chirp mass, generating a
$\delta t_\mathrm{gw}>1$ns. All the black elements refer to our default disk model,
the red elements are for a GW driven MBHB population. Distributions are averaged over $N_k=1000$ realizations
of the MBHB population.
}
\label{f:N(t_gw)distribution}
\end{figure*}
In Figure~\ref{f:N(t_gw)distribution} we plot the distribution of the number of sources (total and resolvable) as a function of
timing residual, detection frequency, redshift, and chirp mass,
found in two particular realizations of our default $\alpha$-disk (all mergers wet) and in the purely GW-driven models (all mergers dry).
The figure shows that even though there are much more sources in the purely GW-driven case, the number of sources rising above
the stochastic level is almost the same in the purely dry and wet cases. Figure \ref{f:N(t_gw)distribution}
also shows the chirp mass, redshift, and frequency distribution for sources above 1\,ns timing level.
As was previously shown in SVV09, the bulk of the sources
are cosmologically nearby ($z \lower .75ex\hbox{$\sim$} \llap{\raise .27ex \hbox{$<$}} 1$) with masses peaking around ${\cal M}\sim 2\times 10^8\,{\rm M_\odot}$.
Figure \ref{f:N(t_gw)distribution} shows that gas dynamics does not introduce
a major systematic change in the shape of the redshift and the chirp mass distributions.
The lower left panel shows that gas removes a systematically larger fraction of sources at
small frequencies.
\begin{figure}
\centering
\mbox{\includegraphics[width=84mm]{fig5.eps}}
\caption{Cumulative number of total ($N_t(\delta t_{\rm gw})$, thin lines) and individually resolvable
($N_r(\delta t_{\rm gw})$, thick lines) sources emitting above a
given $\delta {t_{\rm gw}}$ threshold as a function of
$\delta {t_{\rm gw}}$. {\it Upper panel}: $\alpha$--disk model with $\dot{m}=0.3$.
Lines are for $\alpha=$0.3 (solid), 0.1 (long--dashed) and 0.01 (short--dashed).
{\it Central panel}: same as upper panel but for a $\beta$-disk model.
{\it Lower panel}: nonsteady IPP model, assuming $\alpha=0.3$ and $\dot{m}=0.3$
(solid) and 0.1 (long--dashed). In all the panels, the dotted lines refer,
for comparison, to the GW driven model.
All the distributions refer to the ensemble mean computed over all $N_k=1000$
realizations of the MBHB population.}
\label{f:N_resolvable(t_gw)}
\end{figure}
\begin{figure}
\centering
\mbox{\includegraphics[width=84mm]{fig6.eps}}
\caption{SNR distribution of individually resolvable sources. Thin and thick curves refer
to neglecting the instrumental noise or using a $\delta t_{{\rm gw}}=1\,$ns timing precision, respectively.
Linestyle as in the upper panel of Figure~\ref{f:N_resolvable(t_gw)}.}
\label{f6}
\end{figure}
Figure~\ref{f:N_resolvable(t_gw)} shows the cumulative number of binaries (total and resolvable) as a function of
timing residual. The upper panel shows results for the $\alpha$--disk
models with $\dot{m}=0.3$ and different $\alpha$. The statistics of resolvable
sources is almost unaffected by the large suppression of the total number of sources at a fixed timing residual. For example,
in all of our models, we expect $\sim 2$ resolvable sources at a timing level of $\delta t_{{\rm gw}} = 10$\,ns, even though the total number of
sources contributing to the signal at that level spans about an order of magnitude among the different models
($\sim 5$ for $\alpha=0.3$ to $\sim 50$ for binaries driven by GW only). The same is true for $\beta$--disk models
{ (central panel) and for nonsteady models (lower panel)},
even though in these cases the total number of sources at a particular $\delta t_{{\rm gw}}$ is not reduced dramatically by gas effects.
This result can be understood with a closer inspection of Figure~\ref{figtres}. Let us focus on the $\alpha$--disk model.
As explained in Sec. 3.2, the impact of gas driven dynamics in the PTA window is more significant for lower binary masses
and unequal mass ratios. These
are the binaries that build-up the bulk of the signal, and its stochastic level is consequently
greatly reduced in the gas driven case. On the other hand, the population of high equal mass binaries, which constitute most of the
individually resolvable sources,
is almost unaffected
by the presence of the circumbinary disk, as they are already in the GW-driven regime in the relevant range of binary periods.
Figure~\ref{f6} shows expectations on the detection significance of resolvable sources. The
SNR distribution of resolvable sources (equation
[\ref{e:snr}]) are shown as thin lines, accounting for the astrophysical GW noise from the unresolved binaries,
but neglecting the intrinsic noise of the array (i.e. assuming an ideal detector with infinite
sensitivity, $\delta t_\mathrm{p}^2(f)=0$, in equation \ref{e:noise}). This calculation represents
{\it an upper limit} of the SNR. Thick lines, in Figure~\ref{f6} plot
the SNR considering a total detector noise of 1ns (appropriate for SKA).
The figure shows that the expected detection significance of resolvable sources is systematical higher for gas driven models with larger $\alpha$.
In general, in an array with 1ns sensitivity, we may expect a couple of individually resolvable sources with $\rm SNR>5$.
For near future instruments with much worse sensitivity, the identification of resolvable sources might be
more challenging.
\subsection{Stochastic background}
\begin{figure}
\centering
\mbox{\includegraphics[width=84mm]{fig7.eps}}
\caption{Influence of the gas driven dynamics on $h_c$ (thin lines) and on
$h_s$ (thick lines). {\it Upper left panel}: GW driven dynamics versus gas driven dynamics for our default disk model.
{\it Upper right panel}: $\alpha=$ 0.3 (solid), 0.1 (long--dashed),
0.01 (short--dashed), for an $\alpha$--disk with $\dot{m}=0.3$.
{\it Lower left panel}: same as the upper right panel, considering an $\alpha$--disk with $\dot{m}=0.1$.
{\it Lower right panel}: same as the upper right panel, considering a $\beta$--disk with $\dot{m}=0.3$.
The two dotted lines in each panel represent the sensitivity of the complete PPTA survey and an indicative
sensitivity of 1ns for the SKA.
}
\label{f:stochasticRMS}
\end{figure}
Figure \ref{f:stochasticRMS} shows the RMS stochastic level (i.e. after subtracting off the individually resolvable sources,
see Sec.~\ref{s:statistics}) for selected steady state gas disk models, averaged over $N_k=1000$ realizations.
The top left panel highlights the difference between wet and dry models, the top right and bottom left panels collect different
$\alpha$--disk models and the lower right panel is for selected $\beta$--disk models
{(nonsteady models, not shown here, give results basically identical to $\beta$--disk models)}.
The different line styles show the effect of changing the $\alpha$ parameter of the disk.
We also show the RMS total signal level, which is exactly proportional to $f^{-2/3}$ in the dry case. In general,
the stochastic level matches the total signal level at low frequencies, but is increasingly suppressed for
frequencies above $f \lower .75ex \hbox{$\sim$} \llap{\raise .27ex \hbox{$>$}} 10^{-8}\,\mathrm{Hz}$ for all of our models.
At sufficiently large frequencies GW emission dominates even for wet mergers, and
both the RMS total signal and the stochastic level approaches the purely GW--driven case.
However, at small frequencies, a significant fraction of binaries is driven by gas,
and the signal is attenuated and the spectrum is less steep
compared to the dry case. Figure \ref{f:stochasticRMS} shows that gas--driven
migration suppresses the stochastic background significantly, by a factor of 5 for
our standard disk model below $10^{-8}\,\mathrm{Hz}$.
The suppression of the total and stochastic levels is a strong function of the model parameters.
Interestingly, there is almost no suppression for $\beta$--disks, or for $\alpha$--disks with a
small accretion rate and/or a small $\alpha$ value. In these cases, the local disk mass is smaller,
resulting in longer viscous timescales, and hence the population of widely separated binaries
is not reduced significantly.
\begin{figure}
\centering
\mbox{\includegraphics[width=84mm]{fig8.eps}}
\caption{Variance of the expected stochastic level of the signal as a function of $\alpha$. In all
the panels we assume the $\alpha$--disk model with $\dot{m}=0.3$ (except for the lower right panel,
referring to the GW driven model). Solid lines represent the median $h_s$ over 1000 Monte Carlo realizations,
while the shaded area enclosed within the two dashed lines is the $10\%$--$90\%$ confidence region . The two dotted
lines in each panel represent the sensitivity of the complete PPTA survey and an indicative sensitivity of 1ns for the SKA.
}
\label{f:stochasticVAR}
\end{figure}
Figure~\ref{f:stochasticVAR} quantifies how the stochastic background changes among different
Monte Carlo realizations, showing the range attained by 10--$90\%$ of all $N_k=1000$ realizations.
The variance is not the product of uncertainties related to the parameters of
the adopted cosmological and dynamical model, but is purely determined by the small number statistics
of sources per frequency bin, intrinsic to
the source distribution. At $f=10^{-8}$Hz, the variance of the signal produced by our default disk model
is $\sim 0.5$dex. Decreasing $\alpha$ to 0.1 and 0.01, the variance drops to $\sim0.35$dex and $\sim0.25$dex
respectively. In the case of all mergers driven by GWs, the variance at the same frequency is only $\sim0.15$dex.
This means that, contrary to the GW driven model (SVC08), it is impossible to predict the stochastic level of the signal
accurately for our default gas driven model. For a linear fit, any power law in the range $f^{-0.3}-f^{-1.1}$
is acceptable within the level of variance of the signal in the frequency range $3-30$\,nHz .
\section{Discussion and conclusions}
The presence of a strong nHz gravitational wave signal from a cosmological population of massive
black hole binaries, is a clear prediction of hierarchical models of structure formation, where
galaxy evolution proceeds through a sequence of merger events. The detailed nature of the signal depends
however on a number of uncertain factors: the MBHB mass function, the cosmological merger rate,
the detailed evolution of binaries,
and so on. In particular, MBHB dynamics determines the number of
sources emitting at any given frequency, and thus the overall shape and strength of the signal. Previous works
on the subject (e.g. SVC08 and SVV09) considered the case of GW driven binaries only.
In this paper we studied the impact of gas driven massive black hole binary dynamics on the nHz
gravitational wave signal detectable with pulsar timing arrays. This is relevant because in any
merger event, cold gas is efficiently funnelled toward the centre of the merger remnant, providing a
large supply of gas to the MBHB formed following the galaxy interaction.
Even a percent of the galaxy mass in cold gas funnelled toward the centre, is much larger than the masses
of the putative MBHs involved in the merger so that the newly--formed binary evolution may be driven
by gas until few thousand years before final coalescence.
To conduct our study, we coupled
models for gas driven inspirals of HKM09 to the MBHB population models presented
in SVV09. Our simulations cover a large variety of steady--state and quasistationary one-zone disk models for the
MBHB-disk dynamical interactions (with an extensive exploration of the $\alpha$ viscosity parameter and $\dot{m}$
for $\alpha$ and $\beta$-disks), along with several different prescriptions
for the merging MBHB population (four different black hole mass--galaxy bulge relations coupled with
three different accretion recipes).
The differences with respect to the purely GW driven models (presented
in SVV09) are qualitatively similar for all the considered MBHB populations, we thus presented the results
for the Tu-SA population model only, focusing on the impact of the different disk models.
Our main findings can be summarized as follows:
\begin{itemize}
\item The effect of gas driven dynamics may or may not be important depending on the properties of the
circumbinary disk. A robust result is that if the viscosity is proportional to the gas
pressure only ($\beta$-disk models), there is basically no effect on the GW signal, independently of the other disk parameters.
{ Similarly, there is a tiny effect for time-dependent migration models \citep[i.e.][]{ivanov}, where gas piles up and the accretion rate decreases as the binary hardens.
However, if the viscosity is proportional to the total (gas+radiation) pressure \citep[$\alpha$-disk,][]{ss73},
then the GW signal can be significantly affected for certain steady state circumbinary disk models
\citep{syerclarke} with $\alpha \lower .75ex \hbox{$\sim$} \llap{\raise .27ex \hbox{$>$}} 0.1$ and $\dot{m} \lower .75ex \hbox{$\sim$} \llap{\raise .27ex \hbox{$>$}} 0.3$.}
This difference is explained by the fact that the gas--driven inspiral rate is determined by the radial gas inflow velocity, which is significantly faster for radiation pressure dominated $\alpha$-disks for a fixed accretion rate, as they are more dilute. Therefore gas effects can dominate over the GW--driven inspiral rate for $\alpha$ disks at relatively small binary separations corresponding to the PTA frequency band.
\item With respect to the GW driven case, the presence of massive circumbinary disks affects the
population of low--unequal mass binaries predominantly ($M<10^8\,{\rm M_\odot}$, $q<0.1$), causing a significant suppression of the
{\it stochastic level} of the signal, but leaving the number and strength of massive {\it individually
resolvable sources} basically unaffected. In our default model ($\alpha=0.3$, $\dot{m}=0.3$), the stochastic background
is suppressed by a factor of $\sim 5$ at $f<10^{-8}$Hz. This suppression factor {decreases for smaller $\dot{m}$
and $\alpha$. The stochastic level is {\it not} suppressed significantly for nonsteady disks, for $\beta$--disks (arbitrary $\dot{m}$ and $\alpha$), and for $\alpha$-disks with either $\dot{m} \lower .75ex\hbox{$\sim$} \llap{\raise .27ex \hbox{$<$}} 0.1$ and arbitrary $\alpha$, or $\alpha \lower .75ex\hbox{$\sim$} \llap{\raise .27ex \hbox{$<$}} 0.01$ with arbitrary $\dot{m}$.}
{\it About 10 individual sources are resolvable at 1\,ns timing level, independently of the
adopted disk model.}
\item All the results shown here for the Tu-SA model, hold for every other MBHB population model we tested.
There is a certain level of degeneracy between disk dynamics and MBHB mass function: the stochastic level
given by a population of heavy binaries evolving by gas dynamics, can mimic that of
a population of lighter binaries that are driven by GWs only. However, the variance
of the signal would be much bigger in the former case, because the signal is produced by fewer massive sources.
\item The detection of GWs emitted by MBHBs embedded in gaseous disks with high viscosity and accretion rate,
may be very challenging for relatively short term PTA campaigns like the PPTA. In fact, we find that most of the 12 MBHB population
models tested in SVV09 would {\it not} produce a stochastic signal detectable by the PPTA (i.e. the signal is
a factor of three below the PPTA capabilities for the Tu-SA model). However, long term projects like the
SKA, which aim to nanosecond sensitivities, are expected to be able to detect the GW signal, resolving a handful of
individual sources with high significance.
\end{itemize}
A word of caution should be spent to stress the fact that our models are idealized in many ways. We considered
circular binaries only. If most systems were significantly eccentric, then the overall signal would be modified
by the multi-harmonic emission of each individual source.
Moreover, we only considered radiatively efficient, geometrically thin, one-zone, { steady--state and quasistationary
accretion disks.
The most massive but gas driven binaries, emitting at low GW frequencies $f_{\rm gw} \lower .75ex\hbox{$\sim$} \llap{\raise .27ex \hbox{$<$}} 5\,$nHz, are
radiation pressure dominated, but the migration rate estimates had to be extrapolated using the
scaling exponents for gas pressure dominated disks.
For even wider massive binaries ($f_{\rm gw}$ near the low frequency observation limit)
the disks are marginally Toomre unstable.
We used simple models for the binary disk interaction, scaling the Type-II planetary migration formulae to
MBHBs. The results are sensitive to the models: the steady--state models of \citet{syerclarke} lead to a
decrease in the stochastic nHz background, while the more sophisticated time-dependent models of
\citet{ivanov} as well as the \citep{Hayasaki09,Hayasaki10} models
lead to almost no effect (see also Sec.~\ref{s:gasdetails}, for a list of caveats). }
Further studies should examine the accuracy of these approximations for gas driven migration in circumbinary disks around MBHBs.
Finally, it is likely that not all the binaries are
gas driven on their way to the coalescence, and the efficiency of the disk-binary coupling may vary from merger to
merger depending on the environmental conditions, which may modify the properties of the expected signal as well.
Nonetheless, our calculations provide clear predictions for the possible attenuation of the stochastic GW background,
which may be confirmed or discarded by ongoing and forthcoming pulsar timing arrays.
\section*{Acknowledgments}
A.S. is grateful to M. `penguin' Giustini and V. Cappa for enlightening discussions.
B.K. acknowledges support by NASA through Einstein Postdoctoral Fellowship grant number PF9-00063
awarded by the Chandra X-ray Center, which is operated by the Smithsonian Astrophysical Observatory for
NASA under contract NAS8-03060, and partial support by OTKA grant 68228.
|
\section{Introduction}
Type Ia supernovae (SNe~Ia) have become an essential tool of
observational cosmology \citep{Riess98b, Perlmutter99, Astier06, Wood-Vasey07, Riess07, Freedman09, Kessler09}. By
studying the distance-redshift relation of a large number of
SNe~Ia over a wide range in redshift, the equation of state of dark energy
can be measured. Although SNe Ia can be calibrated to be good standard
candles, they are affected by systematic effects such as
extinction by dust and gravitational lensing.
Because of gravitational
lensing, most supernovae will be slightly demagnified and some will be
significantly magnified due to the mass inhomogeneities along the line
of sight.
This causes an additional
dispersion in the observed magnitudes and thus in the Hubble diagram. The effect
is greater at high redshift, but is still less than the scatter in the Hubble diagram for
a redshift up to 1, as we see in this paper.
Magnification of SNe Ia can be estimated in 2 ways.
The Hubble diagram residuals, computed assuming the best fit cosmological model, give
an indirect measure of the SN
magnification. The magnification can, on the other hand, also be estimated by
modeling the foreground galaxy mass distribution using galaxy
photometric measurements together with derived mass-luminosity
relations for galaxies and dark matter halo models. If a correlation
between these two estimates is found, it could give interesting insight in
the modeling of the mass distribution of the foreground galaxies. This will be discussed in another paper \citep{Jonsson09}
The lensing signature in supernovae samples has already been
sought for. \cite{Williams04}
correlated the brightness of high-z supernovae from the High-z
Supernova Search Team and the Supernova Cosmology Project with the
density of the foreground galaxies. They found that brighter
supernovae preferentially lay behind over-dense regions at a 99 \% CL.
However, they find a lensing contribution to the scatter in the Hubble
diagram of 0.11 mag (for a median redshift below 0.5) which is
larger than expected (for example from lensing statistics), as noted by the authors themselves.
\cite{Menard05} remark that the measured size
of the effect is also incompatible with shear variance measurements.
Finally, such a large magnification effect would cause a significant
increase of Hubble diagram scatter with redshift, undetected today.
\cite{Wang05} derived
the expected weak lensing signatures of Type Ia SNe by convolving the
intrinsic distribution in peak luminosity with expected magnification
distributions and compared the expected and measured
skewnesses of the residual distribution of 110 high and low-z SNe
from the Riess sample \citep{Riess04}. The signal found is dominated
by high redshift events uncorrected for the NICMOS non-linearity.
Its origin remains unknown and its significance is not provided.
Partly using the same supernova sample, \cite{Menard05}
searched for a correlation using SDSS photometry of the foreground
galaxies and found no evidence for a correlation.
More recently \cite{Jonsson06} found a weak correlation with a confidence
level of $90\%$ using high-z supernovae from the GOODS fields.
A firm detection of this effect therefore remains to be found.
The SNLS sample is currently the
best suited sample for a possible detection of such a signal thanks to its
large number of SNe~Ia at high redshift \citep{Jonsson08}.
In this paper, we investigate the possible correlation between the
magnification of the SNLS SNe derived from photometric measurements of
the foreground galaxies (together with chosen mass-luminosity relations)
and the residuals from the Hubble diagram.
Evaluating the expected magnification by foreground galaxies can be
schematically split into three steps : first obtaining redshifts
and absolute magnitudes of the foreground galaxies, then converting
these informations into mass, and finally turning this
set of mass and redshift estimates into a magnification estimate.
The outline of the paper is as follows.
$\S$\ref{sec:mass-luminosity-relations} introduces the mass-luminosity
relations used in this analysis.
In $\S$\ref{sec:data} the SNLS 3-year data set
is presented. $\S$\ref{sec:estimating-magnifications}
describes the analysis steps to estimate the supernovae magnifications,
while in $\S$\ref{sec:lensing-signature} we present the measured correlation
and compare it with expectations from simulations.
Finally, we summarize in $\S$\ref{sec:conclusion}.
\section{Mass Luminosity relations}
\label{sec:mass-luminosity-relations}
There are several ways of inferring the combined mass of a galaxy and its dark matter halo using the measured luminosity.
We make use of two approaches in this paper.
i) The line-of-sight velocity dispersions in elliptical galaxies and rotational velocities
in spiral galaxies are correlated with their luminosities. Those are the well established Faber-Jackson \citep{Poveda1961, Fish1964, Faber1976} and Tully-Fisher \citep{Tully1977, Haynes1999} relations .
ii) Galaxy-galaxy lensing measurements allow a relation to be determined between luminosity and mass once a functional form is assumed for the mass density profile.
We present below recent results for these approaches and compare them assuming a Singular Isothermal Sphere dark matter density profile (hereafter SIS) which has proven to be a good fit
to lensing galaxies \citep{Koopmans06}.
\subsection{Tully-Fisher relation for spiral galaxies}
\label{sec:tully-fisher}
We have chosen to use the \cite{Boehm04} results for
the Tully-Fisher (TF) relation. The maximum rotational velocities of 77
spiral galaxies in a redshift range of
$0.1 < z < 1.0$ were measured in the FORS Deep Field. Anchoring the TF relation at low $z$ using the results from \cite{Pierce1992}, they obtain the
following relation between the maximum rotation velocity $v_{\rm max}$
and the absolute magnitude $M_B$ of the galaxy in the rest-frame $B$-band (in a Vega magnitude system):
\begin{equation}
\log v_{\rm max}=-0.134\left(M_{B}+(3.61\pm0.24)+(1.22\pm0.56)\cdot z\right)
\label{eq:tully-fisher}
\end{equation}
The observed scatter of $M_{B}$ about this relation is $0.41$ mag (r.m.s.).
The redshift dependence expresses the fact that a galaxy of a given mass
was brighter in the past as it hosted a younger stellar population \citep{Barden03,Milvang-Jensen03, Bamford06, chiu08}.
The maximum rotational velocity $v_{\rm max}$ of spiral galaxies, measured at large galactic radii, is dominated by the dark matter mass. Assuming a SIS dark matter density profile for which the rotational velocity is constant, we have $v_{\rm max} \simeq v_{\rm rot}(r \rightarrow \infty) = \sqrt{2} \sigma$, where $\sigma$ is the one-dimensional r.m.s. of the velocity distribution.
\subsection{Faber-Jackson relation for elliptical galaxies}
\label{sec:faber-jackson}
We adopt the Faber-Jackson relation (FJ) by \cite{Mitchell05} (hereafter M05) derived for a sample of $\sim30,000$ elliptical galaxies using
SDSS data. The selection criteria for the sample and the estimate of the
velocity dispersion are described in \cite{Bernardi03I}. The observed
velocity dispersion has been determined by analyzing the integrated
spectrum of the whole galaxy and aperture corrected to a standard
effective radius.
M05 find the following relation between the velocity dispersion and
the absolute magnitude $M_{r}$ of the galaxy in the rest-frame $r$-band (in a AB system):
\begin{equation}
<\log(\sigma)>=2.2-0.091(M_{r}+20.79+0.85 z)
\label{eq:faber-jackson}
\end{equation}
As for the TF relation, the redshift dependence accounts for the evolution of
the age and hence luminosity of the stellar population.
The scatter in the FJ relation induces an uncertainty in the estimate of the
velocity dispersion which has been given by \cite{Sheth03}:
\begin{equation}
\mathrm{r.m.s.}(\log\sigma)= 0.79 (1+0.17(M_{r}+21.025+0.85z)) \nonumber
\end{equation}
The measured velocity dispersions are aperture-corrected central velocity
dispersions which are almost equal to the dark matter
velocity dispersions (see \citealt{Franx1993, Kochanek1994}), so we can identify those
measurements with the velocity dispersion parameter of the SIS profile.
We choose to translate the FJ r-band relation into a restframe B-band relation since restframe B is covered with SNLS griz bands up to $z\simeq1$.
To convert the SDSS $r$-band absolute magnitudes in the AB system to standard $B$-band
Vega absolute magnitudes, a typical color $M_{B}-M_{r}$ for
ellipticals in the AB system is estimated yielding $M_{B}-M_{r}=1.20$\footnote{With the galaxy spectral sequence described in $\S$\ref{sec:photometric-redshifts}, and for the range of rest-frame $U\!-\!V$ colors we consider to select photometrically elliptical galaxies, we obtain on average $M_{B}-M_{r}=0.86$ with an r.m.s of 0.07. This corresponds to an average change of $\sigma$ of 7\% with an r.m.s of 1\% (see Eq.~\ref{eq:faber-jackson}). We ignore this small correction in the following.}
\citep{Gunnarsson06} and an AB to Vega relation $B_{\rm AB}=B_{\rm Vega}-0.12$
is adopted.
\subsection{Galaxy-galaxy lensing mass estimates}
The galaxy-galaxy lensing signal manifests itself by images of
background (faint) galaxies being distorted by foreground (brighter)
galaxies \citep{Hoekstra04, Hoekstra05, Parker05, Kleinheinrich04, Mandelbaum06}.
Unfortunately, one can only study ensemble averaged
properties because the weak-lensing distortion induced by an
individual galaxy is too small to be detected. The measured quantity is
the mean tangential shear which is proportional to the total lens
galaxy-halo mass. Since the lensing signal depends on the angular
diameter distance between observer, lens and source,
galaxy-galaxy lensing results strongly depend on the properties of the galaxy sample, which differ for different surveys.
We use the galaxy-galaxy lensing result from \cite{Kleinheinrich04}
(hereafter K06). The data is taken from the COMBO-17 survey which
consists of observations in five broad-band filters (UBVRI) and 12
medium-band filters. They report a mass to luminosity relation for the full
sample probed out to a maximum radius of 150~h$^{-1}$kpc when modeling
the lenses as SIS profiles:
\begin{equation}
\sigma=156^{+18}_{-24}\left(\frac{L}{10^{10}h^{-2}L_{r\odot}}\right)^{0.28^{+0.12}_{-0.09}} \mathrm{km}.\mathrm{s}^{-1}
\label{eq:galaxy-galaxy-lensing}
\end{equation}
where $L$ is the luminosity of the galaxy. The fiducial
luminosity, $L^{*}=10^{10}h^{-2}L_{r\odot}$ is given in the SDSS
$r$-band. For a conversion to the $B$-band, they have calculated that the
galaxies in their sample with a fiducial luminosity of
$L_{*}=10^{10}h^{-2}L_{B\odot}$ in the $B$-band have a fiducial
luminosity of $L_{*}=1.1\times10^{10}h^{-2}L_{r\odot}$ in the SDSS
$r$-band.
Note that for the SIS profile, the halo is characterized by its velocity dispersion which is the key parameter entering in the estimation of the magnification (see section \ref{sec:gravitational-magnification}).
\subsection{Comparison assuming a Singular Isothermal Sphere profile}
In Figure \ref{lum_mass}, we show the velocity dispersion as a
function of absolute $B$-band magnitude for the three different
mass-luminosity relations
(Eq.~\ref{eq:tully-fisher},\ref{eq:faber-jackson},\ref{eq:galaxy-galaxy-lensing}
for TF, FJ relations and galaxy-galaxy lensing results of K06) assuming a SIS
dark matter mass profile. The mean redshift of the lens galaxies in
the K06-relation is $\sim0.4$. For a given $B$-band luminosity,
elliptical galaxies are more massive than spirals so the FJ relation
naturally gives higher velocity dispersion than
the TF relation. The K06 relation and the FJ relation appear quite similar.
For high luminosity galaxies
the difference in mass estimate from the three relations can lead to very different magnifications. As an example, for a bright galaxy
of absolute magnitude M$_{B}$=-22, the velocity
dispersions range from 170~km s$^{-1}$ (TF) to 249~km s$^{-1}$ (FJ) and 248~km s$^{-1}$
(K06).
As a consequence, in order to obtain accurate estimates of the magnification it is important to take into account the galaxy type. \\
\begin{figure}[h]
\begin{center}
\includegraphics[width=1.1\linewidth]{lum_mass_rel}
\caption{Velocity dispersions as a function of the absolute magnitude in the $B$-band for three different mass-luminosity relations
at a redshift of $z=0.4$. In green (dash-dotted) the K06 relation, in red (dotted) the FJ relation and in
black (solid) the TF relation. The plot shows the velocity dispersion in km s$^{-1}$
as a function of absolute magnitude in the $B$-band. }
\label{lum_mass}
\end{center}
\end{figure}
\section{Data}
\label{sec:data}
The SuperNova Legacy Survey consists of an imaging survey which is
part of the (deep component of the) Canada-France-Hawaii Telescope
Legacy Survey\footnote{see http://www.cfht.hawaii.edu/Science/CFHLS} and a spectroscopic survey done on 8m class telescopes. The imaging was done with the MegaCam imager (360
Megapixels, 1~deg$^{2}$) which detects and monitors the light curves of
the SNe, over four different fields of 1~deg$^{2}$ (called D1, D2, D3, D4).
D1 overlaps partly with the VVDS survey\footnote{http://www.oamp.fr/virmos/vvds.htm}, D2 with the COSMOS survey\footnote{http://cosmos.astro.caltech.edu/}, and
D3 with the DEEP2 survey\footnote{http://deep.berkeley.edu/}. We will later
use the spectroscopic galaxy redshifts from the DEEP2 and VVDS surveys
to establish and test photometric galaxy redshifts.
The "rolling" search method was used for the $griz$ bands,
which means observing the same field every 3-4 days during dark and gray time
(typically for a period of $\sim16$ days around new moon)
for as long as it remains visible (from 5 to 7 months a year). Images in $u$ are also used in this analysis.
The images are used both to monitor the SN light curves
and to build photometric catalogs of the objects in the fields, in
particular the galaxies.
For each field, the total exposure times used
to build the galaxy catalogs for this analysis amount to more than 20 h
per band, reaching typically 60 hours in the $i$-band.
Spectroscopy of the supernovae is
crucial in order to obtain
redshifts, and to confirm the type of each supernova. The
spectroscopic follow-up for the SNLS was done with the VLT~\citep{Baumont08,Balland09}, Gemini~\citep{Howell05,Bronder08} and
Keck telescopes~\citep{Ellis08}.
The 3-year data set used in this paper consists of 233
spectroscopically confirmed Type Ia supernovae used for cosmological
analyses. A detailed description of the SN survey
and the SN data analysis methods is provided in \cite{Astier06,Guy09}.
An important feature, relevant here, is that the
SNLS Hubble diagram exhibits a r.m.s scatter of 0.16 mag.
\section{Estimating the supernovae magnification}
\label{sec:estimating-magnifications}
We describe in this section the analysis steps followed to calculate the supernovae magnifications.
We need a catalog of the field
galaxies with an estimate of the redshift and the rest-frame $B$, $V$ and $U$
band absolute magnitudes for each galaxy. Computing the restframe magnitudes requires, in addition to the redshift,
the knowledge of the galaxy SED.
The rest-frame $B$-band absolute
magnitude is used to convert the luminosity of the galaxy into a
mass estimate using the mass-luminosity relations introduced in
section \ref{sec:mass-luminosity-relations} whereas the $U\!-\!V$ color is
used to separate the galaxies into a red and a blue population.
The photometric catalog is described in $\S$\ref{sec:catalog}. The method we have used to determine photometric redshifts
and absolute magnitudes is accounted for in $\S$\ref{sec:photometric-redshifts}, the galaxy photometric classification is presented in $\S$\ref{sec:spiral-elliptical-classification}, and the estimation of the SN magnification is done in
$\S$\ref{sec:gravitational-magnification}.
\subsection{Galaxy photometric catalog}
\label{sec:catalog}
The galaxy catalogs are built on deep image stacks in the $ugriz$
filters. The deep stacks are constructed by selecting 80 \% of
the best quality images (6241 images for the four fields). Transmission
and seeing cuts (e.g. {\sc fwhm} $< 1.15\arcsec$) are applied. Because we
have fewer exposures in the $u$-band than in the others, less
stringent quality cuts are applied to these images. The selected
images are co-added using the {\sc swarp v2.10}
package\footnote{http://terapix.iap.fr/soft/swarp/}. The source
detection and photometry is performed using SExtractor V2.4.4 (\citep{Sex}) in
double image mode. The detection is made in the i band.
A cut on the signal-to-noise ratio
in the i-band, $S/N>15$ has been made so as to maximize the signal detection while
minimizing the many spurious detections around stars halos. We then
use the AUTO SExtractor flux, computed in an elliptic aperture to
extract the galaxies.
The different cuts lead to a limiting magnitude
$i=25$ (Vega magnitude system).
Zero points are computed using circular photometry on a tertiary stars
catalog, described in \citep{Regnault09}.
Two categories of objects
have to be identified in
the catalog : stars, and the host galaxies of the SNe.
In order to identify stars, we estimate the second moments of all objects in the catalog using a 2D Gaussian fit. We then look for an accumulation
in the second moment $m_{xx}-m_{yy}$ space, taking into account the
variations of the PSF across the focal plane. Objects in this clump
are assumed to be stars.
Since the photo-z of the host galaxy is less precise than the spectroscopic
z of the SN it is possible to have the configuration of the host galaxy being
in the line-of-sight of the supernova and hence wrongly contributing to its magnification.
As a consequence it is important to identify and exclude the host galaxy from the analysis.
Note that identifying the host galaxy on the images is not always obvious.
The host galaxy is identified using two criteria: the minimum distance to the
supernova location and the match between the photometric redshift of the galaxy and
the spectroscopic redshift of the supernova.
To estimate these quantities it is necessary to use images uncontaminated by the supernova light.
For a given supernova, we exclude the images
taken during the same season, i.e. the 6 consecutive months during
which the field was observed. The photometric redshift is derived using the
SNLS photometric redshift code (see section \ref{sec:photometric-redshifts}).
A normalized distance $d$ is
computed using SExtractor shape parameters, so that $d<1$ defines the
photometric ellipse of the galaxy. When no object is found within
$d<1.3$ of the SN, the host of the SN is undetected.
When more than one object is detected close to the
supernova location, we check the match between the galaxies
photometric and the supernova spectroscopic redshifts.
Ambiguous cases
are flagged as problematic and can lead to exclusion of the SN if the
uncertainty in the determination of the host galaxy has an important impact
on the magnification of the SN in question. Four SNe from the sample
were initially excluded in this way.
However, we are fortunate to have HST imaging from the COSMOS
field \citep{Scoville07,Koekemoer07} together with newly published high resolution redshifts \citep{Ilbert09}, also from the COSMOS for
galaxies in the D2 field. As a consequence, two of the initially excluded
SNe have been reexamined, SNLS-04D2kr and SNLS-05D2bt. These two
SNe are detected very close to a presumed host galaxy
which photometric redshift does not match the spectroscopic
redshift of the SN.
Figure \ref{04D2kr} and \ref{05D2bt} show the SN location on
the CFHT and the HST images for the two SNe in question.
For SNLS-04D2kr at $z=0.744$, a smaller and hardly visible galaxy is
detected at the location of the SN in the HST image, very close to the
presumed -- large -- host galaxy. The redshift assigned to this large galaxy from \citep{Ilbert06}, \citep{Ilbert09}, and the SNLS photometric redshift code is $z = 0.168$, 0.228 and
0.3 respectively, implying that the large galaxy in question is not the
host galaxy, but a foreground galaxy. The host galaxy is likely to be the
small elongated galaxy detected in the HST image.
The same situation arises for SNLS-05D2bt at\\
$z=0.68$ : the presumed host galaxy has a photometric redshift of $z=0.31$
and $0.32$ from the SNLS photometric redshift code and \citep{Ilbert06}
respectively.
\begin{figure*}[!t]
\begin{minipage}[t]{0.47\textwidth}
\subfigure[CFHT image.]{\includegraphics[width=0.95\linewidth]{04D2kr_cfht}}
\subfigure[HST image from the COSMOS survey]{\includegraphics[width=0.95\linewidth]{04D2kr_hst}}
\caption{Images of the field around the supernova SNLS-04D2kr at $z=0.744$. The images span 30\arcsec and are oriented north-up/east-left. A red circle indicates the supernova position. A blue ellipse indicates the closest galaxy, as detected on the CFHT image.}
\label{04D2kr}
\end{minipage}\hfill
\begin{minipage}[t]{0.47\textwidth}
\subfigure[CFHT image]{\includegraphics[width=0.95\linewidth]{05D2bt_cfht}}
\subfigure[HST image from the COSMOS survey]{\includegraphics[width=0.95\linewidth]{05D2bt_hst}}
\caption{ Field around SNLS-05D2bt at $z=0.68$. As for SNLS-04D2kr, a smaller and fainter galaxy is conspicuous on the SN location on the HST image.}
\label{05D2bt}
\end{minipage}
\end{figure*}
We obtain in this way a photometric catalog in the $ugriz$ bands,
where stars and supernovae host galaxies are identified. The host
galaxy magnitudes are free from supernova light.
\subsubsection{Masks}
Halos and diffraction spikes around bright stars generate spurious galaxy detections in the catalog.
The size of the halos and the shape of the spikes are fixed by the geometry of the telescope optics,
but their intensity in the images is directly proportional to the brightness of the stars.
Also, for the brightest stars, pixels reaching their saturation level induce bleedings in the
CCD images which have to be accounted for.
Due to those effects, circles with radius varying from 50 to 600 pixels (10 to 120\arcsec)
centered around bright stars are masked out. Those masked regions represent 22\% of the field of view.
An example is displayed in Figure \ref{maskfig} for the D1 field.
\begin{figure}[htp]
\begin{center}
\includegraphics[width=0.95\linewidth]{mask_new}
\caption{A section of the D1 field (5000x4500 pixels or 15.4x13.9\arcmin) with the masked areas shown as purple circles.}
\label{maskfig}
\end{center}
\end{figure}
For each supernova we have included galaxies within a radius of 60\arcsec
(see $\S$\ref{sec:gravitational-magnification}). This will be referred to
as the selected area. In theory, one should exclude all
SNe where the selected area overlap a masked area, but this means
rejecting about half of the SNe. However, the effect on the magnification of small
spurious galaxies or diffraction spikes in the outskirts of the selected area
is small and we chose to keep all SNe with an angular separation larger than 40\arcsec of masked areas.
This leads to the exclusion of 53 SNe.
All SNe have also been looked at by eye to see
whether there are other effects that could lead to exclusion,
e.g. non-masked stars very close to the line of sight leading to the
possibility of excluding an important galaxy hidden behind the star or
leading to possible bad photometry for the surrounding
galaxies. Looking at the SNe we exclude 9 of them and as a conclusion
we keep 171 SNe out of 233 in the initial
sample.
\subsection{Photometric redshifts}
\label{sec:photometric-redshifts}
High quality photometric redshifts have been published by
\citet{Ilbert06} for the galaxies in the SNLS fields down to
$i_{\rm AB}\simeq24$. In these catalogs, fluxes of galaxies hosting
supernovae are affected by the supernovae light, leading sometimes to unreliable
typing and redshift for these galaxies. We also need the SED used to derive the photometric redshift, so as to compute
the restframe magnitudes. Finally, we need to be able to propagate easily the
uncertainties of the photometric measurements to the photo-z and the
absolute magnitude estimations. For these reasons, and so as to
control the error propagation path, we have chosen to derive the
photometric redshifts and the absolute magnitudes as follows.
We first define a continuous one parameter ($a_{*}$) galaxy spectral
sequence, $F(a_{*}, \lambda)$. Indeed as we have two other parameters
to estimate (magnitude and redshift), we cannot afford more than a
single parameter to index the diversity of galaxies in order to keep
one constraint, when, even though 5 bands are used, only 4 are well measured ($griz$, $u$ being
significantly shallower)~\footnote{It would be conceivable to add an extinction
parameter if we restricted our scope to bright galaxies.}. To define
this spectral template, we use the galaxy evolution
model PEGASE.2~\citep{Fioc99}, using a variety of
galaxy SFR law $\propto (t/\tau)\times \exp(-t/\tau)$ and galaxy ages. The initially zero gas metallicity
evolves through the successive generations of stars and ranges from $Z\!\sim\!0.$ to $0.03$.
Extinction is computed using a transfer model for an inclination-averaged disk distribution,
the optical depth being estimated from the mass of gas and the metallicity. We left aside the possibility to add an additional extinction.
The next step consists in optimizing the spectral sequence so as to
reproduce the observed colors of our data in the best way.
It can be corrected to about 30 \%, so that the final template spectra do differ from the initial computed template : our technique is thus marginally
sensitive to the input model.
The
training set comprises a sample of galaxies with known spectroscopic
redshift from the DEEP2 survey \citep{Davis2003, Davis2007}. The spectral
sequence consists in the galaxy flux as a function of wavelength and
the age index, and we optimize a multiplicative correction
(a smooth function of wavelength and age index), along with offsets to the
photometric zero points (which can be interpreted as differential
aperture corrections). We find a residual scatter of 0.097, 0.015,
0.035, 0.025, 0.050 mag for $ugriz$ bands respectively.
The performance of the photometric redshift computation is evaluated
using VVDS spectroscopic redshifts available for the D1 field
(3595~galaxies at $0.01<z<1.5$)~\citep{Lefevre2004}. The redshift
residuals (i.e. $\Delta$z=photometric redshift - spectroscopic
redshift) as a function of spectroscopic redshift are shown in
Figure \ref{zspec}. For $i_{\rm AB}<24$, the fraction of catastrophic failures for
which $\Delta z/(1+z)>0.15$ is $6\%$. Eliminating
catastrophic failures, we obtain $\sigma_{\Delta z} = 0.065 $ and $\sigma_{\Delta z/(1+z)} = 0.037$.
Note that we do not apply any prior on the photometric redshift,
as a large fraction
of the SNLS galaxies are much fainter than those of the training set.
\begin{figure}[h]
\begin{center}
\includegraphics[angle=90.,width=1\linewidth]{zphot-z_spec}
\caption{Performance of the photometric redshift determination: the difference $\Delta z$ between spectroscopic and photometric redshift
as a function of spectroscopic redshift. The spectroscopic
redshifts were obtained from the VVDS survey \citep{Lefevre2004}.}
\label{zspec}
\end{center}
\end{figure}
Uncertainties are estimated using a Monte Carlo propagation of
magnitude uncertainties accounting for the measurement uncertainties
and the residual scatter obtained from the training procedure. In
Figure~\ref{rms_vs_z} we compare the scatter of $\Delta z$ for
the VVDS test sample to the uncertainty on the photometric redshift derived from the
Monte Carlo propagation. Both are in reasonable agreement which
validates our method for propagating uncertainties.
\begin{figure}[h]
\begin{center}
\includegraphics[angle=90.,width=1.0\linewidth]{rms_vs_zphot}
\caption{Performance of our photometric redshift determination: the uncertainty in the estimated photo-z using the
spectroscopic VVDS redshift as a function of photo-z (in dotted red)
and the photo-z uncertainty (solid black line)}
\label{rms_vs_z}
\end{center}
\end{figure}
For some of the galaxies it has been possible to obtain spectroscopic
redshifts from VVDS, DEEP-2 and the SNLS spectroscopic program on VLT (some field galaxies were targeted at the same time as the SN primary target during observations with FORS-2 in multi-slit mode). The D2 field overlaps with the COSMOS field and it has thus been possible to use high resolution
photometric redshifts for a large fraction of the galaxies in the D2
field \citep{Ilbert09}.
\subsection{Classification of spiral and elliptical galaxies based on their colors}
\label{sec:spiral-elliptical-classification}
The Tully-Fisher and Faber-Jackson relations are derived for spiral
galaxies and elliptical galaxies respectively and as a consequence it
is necessary to separate the SNLS galaxies into spirals and
ellipticals. Note that this classification is not strictly necessary:
we could blindly apply some sort of average mass-luminosity relation
to all galaxies. There is indeed no typing involved in
our analysis based on galaxy-galaxy lensing mass estimates.
Separating in broad types is not more than
a way to a priori improve the significance of a potential detection.
The SNLS imaging data does not permit a good morphological classification and
we resorted to using a simple color cut to classify our galaxies.
The rest-frame color $U\!-\!V$ is computed giving rise to two well
separated distributions (the red and the blue population). In Figure
\ref{UV} the rest-frame color $U\!-\!V$ for the SNLS galaxies is shown.
For $U\!-\!V > 0.54$, the
galaxy is classified as an elliptical galaxy or else it is classified
as a spiral galaxy.
Using the rest-frame $U\!-\!V$ color as a proxy for galaxy classification is obviously not optimal,
and the effect of miss-classifying the galaxies
will degrade the lensing signal (see $\S$\ref{sec:correlation}).
\begin{figure}[h]
\begin{center}
\includegraphics[angle=90.,width=1.0\linewidth]{UV}
\caption{Distribution of the rest-frame $U\!-\!V$ color for the SNLS galaxies. A cut at $U\!-\!V=0.54$ separates the distribution into red (elliptical) and blue (spiral) galaxies. }
\label{UV}
\end{center}
\end{figure}
To evaluate whether this cut provides a good estimate for galaxy typing we compare our galaxy types to those
published in \cite{Ilbert06} and \cite{Ilbert09}. They correspond to the galaxy template type, as fitted with eventually some additional extinction, using respectively 5 and 30 band magnitude measurements. For $i_{\rm AB}\leqslant24$, we have classified $89.5\%$ of the COSMOS and $95.3\%$ of the CFHTLS elliptical galaxies as ellipticals. However $56.8\%$ and $26.9\%$ of our elliptical galaxies are classified as spiral galaxies in the COSMOS and the CFHTLS respectively, with an additional extinction of $<E(B\!-\!V)> \simeq 0.2$.
In conclusion, we identify mainly all elliptical galaxies, but there is a contamination of our color-selected elliptical sample with red extinct spiral galaxies.
This contamination will dilute the lensing signal : the red spirals misidentified as ellipticals will be attributed
too high a mass, falsely increasing the supernova expected gravitational magnification, and thus decreasing
its correlation with Hubble diagram residuals.
\subsection{Gravitational magnification}
\label{sec:gravitational-magnification}
The density profile of a SIS can be written
\begin{equation}
\rho_{\rm SIS}(r)=\frac{\sigma^{2}}{2\pi r^{2}}
\end{equation}
The total mass of the halo, $m(r)=2\sigma^{2}r$, diverges and as a consequence
we can use a truncation radius, $r_{t}$ to be able to estimate the total mass of the halo. The truncation radius is chosen to be $r_{200}=\sqrt{2}\sigma/10\,H(z)$ which is defined as the radius within which the mean mass density is 200 times the critical density.
In this case, the convergence is given by
\begin{equation}
\kappa_{\rm SIS}(\theta)=\frac{\theta_{E}}{\pi\theta }\arctan \sqrt{\frac{r_{200}^{2}}{\theta^{2}D_{l}^{2}}-1}
\end{equation}
in the thin lens approximation, where $\theta$
is the transverse angle of the image to the lens, and $\theta_{E}=4\pi (\sigma/c)^{2} D_{ls}/D_{s}$ is the Einstein radius. $D_{ls}$,
$D_{s}$ and $D_{l}$ are respectively the angular distances from the lens to
the source, from the observer to the source and from the observer to the lens, and $\sigma$ is the velocity dispersion.
In order to compute the magnification due to several galaxy lenses,
we use the publicly available software Q-LET~\citep{Gunnarsson04},
which uses the multiple lens plane method.
\begin{figure}[h]
\begin{center}
\includegraphics[angle=90,width=\linewidth]{magnification_rms_vs_r}
\caption{r.m.s of the difference of magnifications $\mu(d_{\rm max})-\mu(d_{\rm max}=125\arcsec)$ as a function of $d_{\rm max}$. This simulation (from real lines of sight) is done for a source at a redshift of 1.}
\label{fig:magnification_rms_vs_r}
\end{center}
\end{figure}
We have investigated the effect on the estimated magnification of
truncating the galaxy list around lines of sight. We have considered
for this purpose random lines of sight in the galaxy catalog, and
computed for each of them the magnification for a SN at $z=1$,
considering sequentially an increasing number of galaxies in the
foreground according to their (transverse) angle $\theta$ to the SN.
Figure~\ref{fig:magnification_rms_vs_r} displays the r.m.s of the
difference of magnifications
$\mu(d_{\rm max}<125\arcsec)-\mu(d_{\rm max}=125\arcsec)$ as a function of
$d_{\rm max}$, where $d_{\rm max}$ is the maximum transverse angle of galaxies
used to compute the magnification\footnote{The maximum value of
$125\arcsec$ was limited by the maximum number of lenses that Q-LET
could handle, this has however essentially no impact on this
analysis: the brightest
galaxy of our sample placed at $125\arcsec$ of a supernova line of sight
causes a magnification $\mu_m < 2\ 10^{-5}$.}.
For $d_{\rm max}=60\arcsec$, the average error on the estimated
magnification due to this galaxy selection is of 0.003. This number is
much smaller than the other sources of uncertainties so that we can
safely ignore galaxies at larger angles.
\subsubsection{Normalization of the magnification distribution}
To estimate the magnification using Q-LET, the lensing galaxies have been put on
top of a homogeneously distributed universe leading to computed gravitational
magnifications always greater than 1 compared to such a universe.
As a consequence the magnifications should be normalized in order to force the
average magnification to be equal to 1, as imposed by flux
conservation.
For this purpose, we have considered numerous random lines of sight
(randomly chosen source positions)
using the true galaxy catalog for a range of source redshifts and computed
the magnification for each of them.
The average magnification for each redshift bin (of 0.1) is recorded and subsequently used to normalize
the magnification estimates.
This correction function is calculated for
each field and each mass-luminosity relation.
\subsubsection{Magnification uncertainty calculation}
\label{sec:mag_uncertainty_calculation}
The uncertainties on the magnifications are evaluated with a Monte Carlo
propagation of the magnitude uncertainties of the galaxies from the catalog and the scatter in the mass-luminosity relations.
Magnitude uncertainties include both measurement uncertainties and the residual scatter of the photometric redshift training (see $\S$\ref{sec:photometric-redshifts}).
A fit of the photometric redshift is performed for each Monte Carlo realization of each galaxy, except for those which have much more precise spectroscopic or photometric redshifts from other sources.
For instance, high resolution photometric redshifts are available for galaxies of the COSMOS catalog \citep{Ilbert09}. For those, instead of fitting for the redshift, we considered random realizations of the redshift within the errors given in their paper, and used our code with those fixed redshifts to determine the uncertainties on the absolute magnitudes.
The scatter in the TF and FJ mass-luminosity relations are well defined from the observations (see $\S$\ref{sec:tully-fisher} and $\S$\ref{sec:faber-jackson}). However, there is no such scatter for the mass luminosity relation derived from galaxy-galaxy lensing observations, as this relation is the result of a global fit to a halo model where each galaxy provide little information. Therefore, for this relation we assume the same scatter as for the TF relation.
In Figure~\ref{erreur_relative} we show the uncertainty of the magnification as a function of the magnification obtained with the TF/FJ relations.
We obtain a relative uncertainty on the magnification of 17\%. The most
important source of uncertainty comes from the scatter in the
mass-luminosity relation. The uncertainties due to redshift are small
and represent a relative uncertainty of about 5\%.
\begin{figure}[h]
\begin{center}
\includegraphics[angle=90.,width=1.0\linewidth]{mag_vs_e_Klein1_pez}
\caption{Magnification uncertainty as a function of magnification both expressed in magnitudes. The line is a straight line fit to the data. The relative uncertainty is about 17\%.}
\label{erreur_relative}
\end{center}
\end{figure}
\section{Signature of the lensing effect}
\label{sec:lensing-signature}
\subsection{Signal discriminant}
As a criterion for a lensing signal detection we have chosen to calculate the weighted correlation coefficient.
\begin{equation}
\rho=\frac{\mathit{cov}(\mu_m,r)}{\sqrt{\mathit{var}(\mu_m)\mathit{var}(r)}}
\end{equation}
with $\mu_m = -2.5 \log_{10}{\mu}$, $\mu$ being the gravitational magnification factor, and $r$ is the residual to the Hubble diagram (hereafter Hubble residual).
The weighted covariance of two variables $x$ and $y$, can be written as :
\begin{equation}
\mathit{cov}(x,y)=\frac{\sum wxy}{\sum w}-\bar{x}\bar{y}
\end{equation}
where $w$ is the weight assigned to each datapoint and $\bar{x}$ and $\bar{y}$ are the weighted means.
Using random lines of sight (ie. random source positions in the galaxy catalog), we have found that weighting with
the inverse of the Hubble residual variance is optimal for
the signal detection. Note that any scaling applied to $\mu_{m}$ and
hence any common offset applied to the mass-luminosity relations
will not change the weighted correlation coefficient and thus not change
the detection significance.
\subsection{Expected Signal}
\label{sec:simulations-expectations}
Before presenting the results, it is useful to have an idea of what to
expect. We now have all the tools needed to perform detailed Monte
Carlo simulations using the true galaxy catalog and give precise
predictions both for the SNLS 3-year data set and the full SNLS
sample.
Magnification distributions for supernovae at different redshifts were
generated by calculating the magnification factor for a sample of simulated
supernovae randomly positioned in the true galaxy catalog. In this way it was possible to
simulate the magnification distribution for a sample of supernovae
with a given redshift distribution, chosen as the actual sample
redshift distribution.
To generate a correlated sample, i.e assuming a correlation
between the Hubble residual and the magnification,
we first computed a ``true'' magnification
using the TF/FJ relations applied to galaxies along the line of sight.
We then drew an expected magnification using the uncertainty
described in \ref{sec:mag_uncertainty_calculation}.
The corresponding Hubble residuals were eventually
generated from the expected magnification and a random offset of r.m.s
0.16 mag in accordance with the SNLS data for which the Hubble
residuals exhibit an r.m.s of 0.16 mag. An uncorrelated sample can be
generated by randomly assigning a Hubble residual to each expected magnification.
We first simulated a large number of correlated samples similar to the real one with 171
supernovae. We compared the distribution of the weighted correlation
coefficient of these samples with that of uncorrelated samples
(see Figure \ref{mc_rho}).
\begin{figure}[h]
\begin{center}
\includegraphics[angle=90,width=1.0\linewidth]{rho_mc_v2}
\caption{Simulated distribution (using real lines of sight) of the weighted correlation coefficients for
correlated samples (dotted) and uncorrelated samples (solid),
for the 3-year SNLS sample.
There is
50 \% probability of finding a 2.5 $\sigma$ significance correlation or better ($\rho=0.22$) and 35 \%
probability of detecting a 3 $\sigma$ signal ($\rho=0.25$).}
\label{mc_rho}
\end{center}
\end{figure}
We find that for the current sample there is 50 \% probability of finding a
2.5~$\sigma$ significance correlation or better and 35 \% probability of detecting a 3~$\sigma$ signal.
For the final SNLS sample we expect $\sim$400
spectroscopically confirmed Type Ia supernovae and $\sim$200 photometrically identified Type Ia
supernovae. Due
to masking, about a third of the supernovae will be rejected. If we do
simulations for 400 supernovae with the same parameters as the current
sample we find that there is 80\% probability of detecting a 3 $\sigma$ signal or more
(see Figure \ref{mc_rho_fullsample} ).
\begin{figure}[h]
\begin{center}
\includegraphics[angle=90,width=1.0\linewidth]{rho_mc_fullsample_final}
\caption{Same distributions as Figure \ref{mc_rho}, for the simulated full SNLS sample. There is 80\% probability of detecting a 3 $\sigma$ signal ($\rho=0.22$) or better. }
\label{mc_rho_fullsample}
\end{center}
\end{figure}
Another question to be addressed is how the errors influence the
possibility of a signal detection. The errors on the magnification are
already small. Monte Carlo simulations performed with various
scaled errors of the magnification support that the scatter in the
magnification has
little impact on the signal detection. The
signal detection is highly dominated by the scatter in the SN
Hubble residuals which is not likely to decrease significantly in the near
future. As a consequence improving the signal detection will require better
statistics and, if possible, higher redshift SNe.
\subsection{Results}
\label{sec:results}
\subsubsection{The SNLS supernovae}
\label{sec:snls_supernovae}
The 3-year data release contains 233 spectroscopically confirmed
Type Ia supernovae in the redshift range 0.2-1.05 after quality cuts.
The Hubble residual of the supernova is the difference in distance modulus of
the supernova and the best cosmology fit. The distance modulus is
estimated from the supernova brightness and other parameters
resulting from a fit of the SALT2 model \citep{Guy07}
to the supernova light curves.
The uncertainty of a distance modulus combines three sources: the
photometric measurement uncertainties, the light curve model scatter
(see \cite{Guy07} for a detailed definition), and the so-called intrinsic
scatter which expresses our lack of complete physical understanding of SN Ia.
This intrinsic scatter is chosen so that
the minimum $\chi^2$ of cosmological fits to the Hubble diagram matches
the number of degrees of freedom. This ensures that the quoted uncertainties
of Hubble residuals properly describe their actual scatter.
\subsubsection{Magnification}
\begin{figure}[!h]
\begin{center}
\includegraphics[width=1.\linewidth]{mag_vs_z_Klein1_pez}
\includegraphics[width=1.\linewidth]{mag_vs_z_TFFJ_pez}
\caption{Magnification factor of the SNLS supernovae versus redshift.
Results based on the K06 luminosity-mass relation (top) and on TF and
FJ relations (bottom). Most of the SNe are slightly demagnified whereas some are significantly magnified.}
\label{mag_z}
\end{center}
\end{figure}
\begin{figure}[!h]
\begin{center}
\includegraphics[angle=90, width=1.\linewidth]{mag_Klein1_pez}
\includegraphics[angle=90, width=1.\linewidth]{mag_TFFJ_pez}
\caption{Magnification distribution of the SNLS supernovae. Results based on the K06 luminosity-mass relation (top) and on TF and FJ relations (bottom). The magnification distributions peaks at a value slightly below 1 and presents a high magnification tail.}
\label{mag}
\end{center}
\end{figure}
\begin{table*}[!ht]
\begin{center}
\begin{tabular}{||c|c|c|c||c|c|c|c||}
\hline
\multicolumn{2}{||c|}{} & \multicolumn{2}{|c||}{magnification factor $ \mu$} & \multicolumn{4}{|c||}{Most important lensing galaxies}\\
\hline
SN & z & K06 & TF-FJ & z(galaxy) & d (\arcsec) & $\sigma$ km/s (K06) & $\sigma$ km/s (TF-FJ) \\
\hline
04D1iv & 0.998 & 1.267$\pm$0.034 & 1.199$\pm$0.031 &0.60 & 7.8& 299 & 295 \\
& & & & 0.51& 5.8 & 217 & 154\\
\hline
04D2kr & 0.744 & 1.208$\pm$0.046 & 1.355$\pm$0.182& 0.228&1.5 &119 & 150 \\
\hline
05D2by & 0.891 & 1.135$\pm$0.016 &1.059$\pm$0.035 & 0.66& 1.8& 89 & 50\\
& & & & 0.68& 4.3 & 151 & 167 \\
& & & & 0.44& 2.7 & 88 & 54 \\
\hline
05D2bt & 0.68 & 1.224$\pm0.33$ & 1.113$\pm$0.033&0.31 &0.5 & 99 & 65 \\
\hline
05D3cx & 0.805 & 1.143 $\pm$0.026 & 1.127$\pm$0.035 & 0.38& 6.4& 224 & 243\\
\hline
03D4cx & 0.949 & 1.179$\pm$0.038 & 1.205$\pm$0.054&0.45 &5.5 & 246 & 259 \\
\hline
04D4bq & 0.55 & 1.196$\pm$0.027 & 1.114$\pm$0.028 &0.32 & 4.4&152 & 108 \\
& & & & 0.38& 4.1 & 190 & 138 \\
\hline
05D4cq & 0.702 & 1.168$\pm$0.028 &1.133$\pm$0.026 & 0.28&15.7 & 282 & 298 \\
& & & & 0.43& 2.9 & 106 & 67\\
\hline
\end{tabular}
\end{center}
\caption{The most magnified supernovae and the characteristics of
the galaxies dominating the magnification.}
\label{top_20}
\end{table*}
Figure \ref{mag_z} and \ref{mag} show the magnification of each SN as
a function of redshift and the magnification distribution
respectively. As expected, most SNe are demagnified with respect to a
homogeneous universe and some are significantly magnified. Moreover,
the magnification distribution peaks at a value slightly lower than
one and presents a long magnification tail. Images of galaxies along the
line-of-sight of 3 of the most magnified supernovae (04D1iv, 03D4cx
and 04D4bq) in the 3-year data set are shown in Figure \ref{sn} (see also Figure \ref{04D2kr} \& \ref{05D2bt}).
All 3 SNe have one or several massive galaxies very close
to the line-of-sight located roughly halfway between the SN and us,
causing the high
magnification. The 8 most magnified supernovae
along with information on the most important galaxies
causing the magnification are listed in table \ref{top_20}.
\begin{figure}[!hb]
\begin{center}
\subfigure{
\includegraphics[width=0.95\linewidth]{04D1iv_zoom_img}
}
\subfigure{
\includegraphics[width=0.95\linewidth]{03D4cx_zoom_img}
}
\subfigure{
\includegraphics[width=0.95\linewidth]{04D4bq_zoom_img}
}
\end{center}
\caption{From the top to the bottom, SN 04D1iv at a redshift of 0.998, SN 03D4cx at a redshift of 0.949 and SN 04D4bq at a redshift of 0.550. In red, the SN and its host. In blue the foreground galaxies labeled according to the closest elliptical distance to the SN and in green the background galaxies. All images span 30\arcsec (horizontally) and are oriented north-up/east-left.}
\label{sn}
\end{figure}
\subsubsection{Correlation}
\label{sec:correlation}
We are searching for a correlation between the Hubble residuals from a best fit
cosmological model and the estimated
magnifications of the supernovae based on foreground galaxy modeling. In
Figure \ref{mag_res_k1} we show a plot of the Hubble residuals of the 171 SNe
versus the estimated magnification.
\begin{figure}[!h]
\begin{center}
\includegraphics[angle=90,width=1.\linewidth]{residu_mag_Klein1_pez}
\includegraphics[angle=90,width=1.\linewidth]{residu_mag_TFFJ_pez}
\caption{The Hubble residuals as a function of magnification for the SNLS. Results based on the K06 luminosity-mass relations (top) and on TF and FJ relations (bottom).}
\label{mag_res_k1}
\end{center}
\end{figure}
The weighted correlation coefficient for this sample is $\rho = 0.12$
using the K06 relation and $\rho = 0.18$ using the TF and FJ relations
respectively. To evaluate the strength of the correlation we calculate the
distribution of the weighted correlation coefficient for an
uncorrelated sample and compare it with the obtained value for our
sample (see Figure \ref{rho}). The uncorrelated samples are drawn
by randomly associating Hubble residuals and expected magnifications of the
real sample. The probability of finding a larger weighted correlation
coefficient than the measured one from an uncorrelated sample is 5\%
using the K06 relation and 1\% using the TF and FJ relations, corresponding
to 1.6 and $2.3\,\sigma$ detections respectively.
It is tempting to attribute the stronger detection of the TF/FJ
relation to the fact that this method considers separately elliptical
and spiral galaxies. In order to test the influence of the separation,
we run the analysis again, but with a random galaxy type assignment
(however preserving the measured proportions). We then find that the
significance of the detection drops from 2.3 to 1.4 $\sigma$, and
conclude that the higher significance of our detection using TF/FJ
relations can be attributed to the distinction between spiral and
elliptical galaxies, rather than chance. So, our primary result is the
detection of supernovae lensing using TF/FJ relations at the 99\% CL.
Note that simulations assuming a perfect galaxy typing (see section \ref{sec:simulations-expectations})
give rise to a mean detection level of 2.5 $\sigma$ (see Figure \ref{mc_rho}).
The 2.3 $\sigma$ detection we find is thus close to the optimal
so that further improvements on galaxy typing should not give rise
to a much larger significance of the signal.
In order to test the scale of galaxy mass estimates, we fit the
slope $a$ relating Hubble residuals and expected magnifications
: $<r> = a < \mu_m>$ and find $a=0.65~\pm~0.30$. Hence the data
are consistent with the TF/FJ mass-luminosity relations
at the 1.2 $\sigma$ level, with a precision of 30\%.
\begin{figure}[!h]
\begin{center}
\includegraphics[angle=90,width=1.\linewidth]{rho_real_TFFJ_pez}
\caption{
Distribution
of the weighted correlation coefficient for an
uncorrelated sample when using TF and FJ relations (in black). The red line indicates the value $\rho = 0.18$
obtained for our sample. The measured correlation significance is $2.3 \, \sigma$ (99\% CL).}
\label{rho}
\end{center}
\end{figure}
Using random lines of sight in the real data, we can estimate the
increase of Hubble diagram scatter expected from gravitational
lensing, as a function of redshift. This is shown in Figure
\ref{fig:magnification_rms} for the TF/FJ relations, which can be
roughly described as $\sigma(\mu_m) = 0.08 \times z$. Alternatively, if
we use the value of $a$ derived from a fit of the relation between Hubble residuals and magnifications, we obtain a lower value $\sigma(\mu_m) = (0.05 \pm 0.022) \times z$.
\begin{figure}[!h]
\begin{center}
\includegraphics[angle=90,width=0.95\linewidth]{magnification_rms}
\caption{
Expected magnification scatter from random lines of sight
as a function of the source redshift, using the TF/FJ
mass-luminosity relations.}
\label{fig:magnification_rms}
\end{center}
\end{figure}
\section{Conclusion}
\label{sec:conclusion}
We have calculated the expected magnification of the SNLS 3-year
sample from the foreground galaxy properties and searched for a
correlation with residuals from the Hubble diagram. A correlation is
detected at the 99\% CL, compatible with a slope of 1.
The expected magnifications cause an extra scatter in the Hubble diagram
approximated by $0.08 \times z$ from the TF/FJ relations we used, which becomes $(0.05 \pm 0.022) \times z $
once these TF/FJ relations are calibrated with the supernova data. We show that separating
the galaxy sample into a blue and a red population based on a $U\!-\!V$
color cut increases the significance of the detection. This is due to the fact that the mass estimate
and hence the induced magnification is significantly different for a red and a blue galaxy of same luminosity (the TF and FJ relations).
Simulations also
point to the fact that a signal detection is dominated by the
number of SNe, their redshift distribution and the scatter in the SN
residuals. Reducing the scatter in the estimated magnification by
increasing the photometric redshift precision or reducing the scatter in the TF and FJ relations
have little effect on the probability of a signal detection.
Finally, simulations using the true galaxy catalog show that using the full
SNLS data set ($\sim$400 expected spectroscopically confirmed Type Ia SNe
and $\sim$200 photometrically identified Type Ia SNe) there is 80\%
chance of detecting a $3 \, \sigma$ signal or more. \\
\begin{acknowledgements}
This article is based on the work of a Ph.D. thesis. TK acknowledges
the support of the French ministry of Education and Research
and the Dark Cosmology Centre funded by the Danish
National Research Foundation. We thank O. Ilbert for providing us
with high resolution photometric redshifts in the COSMOS field prior
to publication. We also thank Michel Fioc, Damien Le Borgne and Jens Hjort for useful discussions. This
research has made use of the NASA/ IPAC Infrared Science Archive,
which is operated by the Jet Propulsion Laboratory, California
Institute of Technology, under contract with the National Aeronautics
and Space Administration. The data reduction was carried out
at the IN2P3 computing center in Lyon, France.
\end{acknowledgements}
\bibliographystyle{aa_like_apj}
|
\section{Introduction}
O stars are amongst the most massive and intrinsically luminous stellar
objects found in galaxies. Since they are the only direct way to measure the masses and radii
of stars, binaries are the perfect testbeds for studying the physical properties and evolution of
such stars. Unfortunately, as reported by \cite{bon08}, less
than 20 O stars have accurate ($\leq$10\%) dynamical mass estimates.
Because of the scarcity of known massive, eclipsing, double-lined,
spectroscopic binaries - and hence dynamical mass estimates for stars at
different evolutionary states \citep[e.g.,][]{gie02} - the mass luminosity relation
and theoretical evolutionary tracks of massive stars ($M\geq20M_\odot$) are
currently poorly constrained by observations.
In order to address this shortfall, much effort has been expended to identify
further examples, utilising photometric and spectroscopic observations of
young massive clusters such as the Arches \citep{mar08},
Quintuplet \citep{fig99}, and Westerlund 1 (\citealt{cla05}; \citealt{rit09}). These observations
indicate that the binary fraction is potentially very high\footnote{A large amount of
observations and patience are necessary to determine the true binarity of a sample of
stars in an open cluster as observed in \cite{san08}};
\cite{kob07} inferred it to be $\geq$ 80\% in Cygnus OB2 (Cyg OB2),
while \cite{cla08} estimate the binary fraction of WR stars in
Westerlund 1 to be $\geq$70\%.
Currently the
stars with the highest dynamical mass estimates are the twin WN6ha components
of WR20a within Westerlund 2 (83$M_\odot$+82$M_\odot$;
\citealt{rau04}, \citealt{bon04}) and the newly discovered
WN6ha binary NGC3603-A1 (116$\pm$31$M_\odot$ + 89$\pm$16$M_\odot$;
\citealt{sch08}). This in turn has implications for the
determination of the Initial Mass Function (IMF) for the clusters in question and,
by extension, the empirically determined maximum mass possible for a
star \citep[cf.][]{fig05}. Moreover, massive close binaries are the progenitors
of such diverse energetic phenomena as supernovae, $\gamma$-ray bursts and
X-ray binaries \citep{rib06}. Clearly the properties of the progenitor binary
population must be known to constrain their formation channels.
Cyg OB2 is one of the most massive and richest associations in
the Galaxy. It is $\sim$2 Myr old and 1.8 kpc away \citep{kim07}.
Containing at least 60-70 O-type stars \citep{neg08}, its proximity and
accessibility to optical studies have made it the focus of numerous
observational campaigns to determine the properties of its massive stellar
population \citep[e.g.][]{mas91, kno00, han03, com02, kim07, kim08}.
Cyg OB2 B17 (\citealt{com02}; henceforth B17 and also known as V1827Cyg,
2MASS J20302730+4113253, NSVS 5738756;
$\alpha_{2000}$=20$^{h}$30$^{m}$27.3$^{s}$,
$\delta_{2000}$=+41$^{o}13^{\prime}$25$^{\prime\prime}$; $V$=12.6) is a luminous,
variable member of the Cyg OB2 association. \citet{com02} observed the
system as part of a near-infrared spectroscopic survey and found the spectrum
presented $Br$$\gamma$ emission, confirming it to be an evolved massive star.
Follow up observations were made by \cite{neg08}, who found it
to stand out from the rest of the members due to its variability and strong emission
lines. They classified it as an Ofpe star and suggested it was a strong binary candidate.
This paper reports the first results of an extensive multi-epoch
photometric and spectroscopic observational campaign on the binary candidate B17.
Section~2 gives a description of
the photometric and spectroscopic observations. Photometrically, the system was found
to be variable and we report the analysis of the light curve in Section~3.
More spectroscopic data were obtained permitting preliminary spectral and luminosity
classification of the system; this analysis is shown in Section~4,
along with descriptions of the long and short timescale variations of the spectra.
The light and radial velocity curve modelling is presented in Section~5.
A discussion of the system, including its evolutionary status, is found
in Section~6 and a summary is presented in Section~7.
Note that the central goals of this manuscript are to verify the binary hypothesis and
present a preliminary spectral classification. A full analysis of the system,
consisting of the deconvolution of an expanded
spectral data set and subsequent model atmosphere analysis to
determine the fundamental stellar parameters of the system will be presented in a
future paper (Stroud et al. in prep).
\section{Observations and Data Reduction}
\subsection{Photometry}
The North Sky Variability Survey (NSVS; \citealt{woz04}) is a record of the
sky at declinations higher than $\delta=-38\degr$ over the optical magnitude range 8
to 15.5. It contains light curves of over 14 million objects. The data were taken between
1999 April and 2000 March by the first-generation Robotic Optical Transient Search
Experiment (ROTSE-I) at Los Alamos National Observatory, New Mexico. The telescope
consisted of four unfiltered Canon 200mm telephoto lenses with f/1.8 focal ratio,
each covering $8\fd2\times8\fd2$. These were equipped with AP10-cameras and
Thomson TH7899M CCDs. The lenses had a typical point-spread function with a full width
half maximum of $\sim$20$^{\prime\prime}$. In a median field, the bright unsaturated
stars had a point-to-point photometric scatter of $\sim$0.02 mag and position errors
within 2$^{\prime\prime}$. The calibrated images were passed through SExtractor software
\citep{ber96}, reducing them to object lists. The data were accessed
through the Sky Database for Objects in Time-Domain (SkyDOT) at Los Alamos National
Laboratory. A total of 186 observations were obtained for B17.
Additional $V$ band photometry was obtained by amateur astronomers Pedro Pastor Seva
(observer 1) and Manuel M\'{e}ndez Marmolejo (observer 2) between 2007 April 03 and 2007 June 11.
Observatory 1 is located in Muchamiel (Alicante, Spain). The telescope used was an 8
inch Vixen VISAC Schmidt-Cassegrain telescope. It has a field of view of
24$^{\prime}$x20$^{\prime}$ and a focal ratio of f/8.The telescope was equipped with a SBIG ST10-XME
CCD chip. Observatory 2 is located in Rota (C\'{a}diz, Spain). The telescope used was an 8 inch
Meade LX200 Schmidt-Cassegrain telescope. It has a field of view of 16$^{\prime}$x12$^{\prime}$
(as a smaller SBIG ST7-XME chip was used) and a focal ratio of f/6.3. Exposures varied between 3 and 4
minutes giving a SNR $\sim$200. The data were calibrated and analysed
using Mira Pro\footnote{Mira Pro software is published by Mirametrics Inc., which has
no connection with the Monterey Institute for Research Astronomy} and
AIP4WINv2\footnote{AIP4WINv2 is published by Willmann-Bell, Inc.} packages. The apparent magnitudes were obtained
using differential photometry with respect to a set of reference stars in the image with known magnitudes.
In both cases the precission for the individual meassures is always better than 0.01 mag.
\subsection{Spectroscopy}
The spectra were obtained from several telescopes during the course of $\sim$three years.
Table~\ref{obs} lists the full set of observations. The first set was observed with the
1.52-m G. D. Cassini telescope at the Loiano Observatory (Italy) during the night
of 2004 July 18. The telescope was equipped with the Bologna Faint Object
Spectrograph and Camera (BFOSC) and an EEV camera. Grism 3 was used, which
covers 3300-5800$\AA$ with a resolution of $\sim$6\AA.
More spectra of the system were obtained with the 4.2-m William Herschel Telescope
(WHT), in La Palma (Spain) equipped with the ISIS double-beam spectrograph, during
service runs on 2006 June 11 and August 18. The instrument was fitted with the R300R grating
and MARCONI2 CCD in the red arm and the R300B grating and EEV12 CCD in the blue
arm. Both configurations result in a nominal dispersion of 0.85\AA/pixel (the
resolution element is approximately 3 pixels in the blue and 2 pixels in the red).
The system was also observed with the 2.5m Isaac Newton Telescope (INT) in La
Palma on 2006 September 10-11. The telescope was
equipped with the Intermediate Dispersion Spectrograph (IDS), fitted with a R632V
grating and EEV10 CCD.
Another 12 spectra were obtained during a dedicated run on 2007 August 21-22 at the WHT.
The system was observed in the blue arm with grating R1200B
(nominal dispersion of $\sim$0.23\AA/pixel).
All the spectra were reduced with the {\em Starlink} packages {\sc ccdpack} \citep{dra00}
and \textsc{figaro} \citep{sho97} and analysed using {\sc figaro} and {\sc dipso} \citep{how98}.
\begin{table}
\caption {Log of Spectroscopic Observations. }
\label{obs}
\centering
\begin{tabular}{c c c c c c c}
\hline\hline
Date &Telescope &Nominal &Wavelength &Phase\\
& /Instrument & Dispersion & Range (\AA) & \\
& & (\AA/pix) & &\\
\hline
04/07/18 &Cassini/BFOSC & 6&3800-6400&0.437
\\
04/07/18 &Cassini/BFOSC & 6&6100-8200&0.450
\\
06/06/11 &WHT/ISISB & 0.86 &3200-5200 & 0.036
\\
06/06/11 &WHT/ISISR & 0.93 &5400-8000 & 0.036
\\
06/08/18 &WHT/ISISB & 0.86 &3200-5200 & 0.929
\\
06/08/18 &WHT/ISISR & 0.93 &5400-8100 & 0.931
\\
06/09/10 &INT/IDS & 0.9 &3900-5000 & 0.600
\\
06/09/11 &INT/IDS & 0.9 &3900-5000 & 0.840
\\
06/09/11 &INT/IDS & 0.9 &3900-5000 & 0.891
\\
07/07/21 &WHT/ISISB & 0.86 &3600-5300 & 0.738
\\
07/07/21 &WHT/ISISR & 0.93 &5400-8300 & 0.738
\\
07/07/21 &WHT/ISISR & 0.93 &5400-8300 & 0.740
\\
07/07/21 &WHT/ISISR & 0.93 &5400-8300 & 0.741
\\
07/08/21 &WHT/ISISB & 0.23 &3900-4750 & 0.380
\\
07/08/21 &WHT/ISISB & 0.23 &3900-4750 & 0.385
\\
07/08/21 &WHT/ISISB & 0.23 &3900-4750 & 0.397
\\
07/08/21 &WHT/ISISB & 0.23 &3900-4750 & 0.416
\\
07/08/21 &WHT/ISISB & 0.23 &3900-4750 & 0.436
\\
07/08/21 &WHT/ISISB & 0.23 &3900-4750 & 0.453
\\
07/08/22 &WHT/ISISB & 0.23 &3900-4750 & 0.629
\\
07/08/22 &WHT/ISISB & 0.23 &3900-4750 & 0.634
\\
07/08/22 &WHT/ISISB & 0.23 &3900-4750 & 0.648
\\
07/08/22 &WHT/ISISB & 0.23 &3900-4750 & 0.664
\\
07/08/22 &WHT/ISISB & 0.23 &3900-4750 & 0.683
\\
07/08/22 &WHT/ISISB & 0.23 &3900-4750 & 0.701
\\
\hline
\end{tabular}
\end{table}
\section{Photometry}
The {\em Starlink} software {\sc period} \citep{dhi01} was used on both the
NSVS and amateur photometry to search for a modulation period in the photometric data,
using phase dispersion minimisation, $\chi^{2}$ of sine fit vs frequency and
string-length vs frequency methods. The resultant periods were consistent and favoured
an orbital period of $4.0217\pm0.0004$ days
(the error being estimated from the spread in the periods derived from the different
methods employed). When both data sets were folded together they were found to be
in phase; therefore, given that there appears to be no change of shape or shift in
period in the 7 years between the NSVS and amateur observations, we are confident
that the period determined is accurate to within the errors quoted.
\begin{figure}[h!]
\centering
\resizebox{\columnwidth}{!}{\includegraphics[trim = 0mm 0mm 5mm 60mm, clip]{12123f1a.ps}}
\resizebox{\columnwidth}{!}{\includegraphics[trim = 0mm 0mm 5mm 60mm, clip]{12123f1b.ps}}
\caption{Upper panel: Periodogram for B17 data using the reduced-$\chi^2$ technique.
Lower panel: Light curve folded on a 4.0217 day period. The red triangles show the phases for which spectra were obtained.}
\label{fig:periodogram}
\end{figure}
Fig.~\ref{fig:periodogram} shows the periodogram obtained using Period with the reduced-$\chi^2$
technique (top) and the
$V$ band light curve folded on a 4.0217 day period, along with the phases of the spectra obtained (bottom).
The shape of the light curve
suggests that the system is a semi-detached binary; Both minima
are narrow and demonstrate different eclipse depths, although the $0.5-0.6$~mag range indicates the
almost complete eclipse of each star
and hence that the two stars are of similar size but with
different luminosities (in a contact system, the temperature of both stars should be the same)
There appears to be an asymmetry on the light curve between
phases $\phi=0.6 - 0.75$ which is observed in both sets of data. \cite{hil05} found similar depressions
in eclipsing binaries in the Small Magellanic Cloud and
\cite{bon08} found a similar depression in the light curve for the binary LMC-SC1-105. They attributed
these asymmetries to the presence of a mass-transfer stream. \cite{lin09} also observed this asymmetry
in the overcontact binary Cyg OB2 $\#$5 and attributed this to the secondary being brighter on its leading
side due to the colliding wind region; first discussed by \cite{rau99}.
The quadratures display an O'Connell effect - where the maxima are of different brightnesses - of $\sim$0.02 mag which
is likely due to variations in the brightness of the stellar surface(s) due to mass transfer between the components.
The principal minimum was estimated using the amateur data to be at $JD=2,454,272.527\pm0.005$.
From the light curve, the ephemeris for the primary eclipse was found to be:
\begin{equation}
minI~=~2454272.527~+~4.0217E~~{\rm (JD)}
\end{equation}
\noindent where $E$ is the number of orbital cycles after the given epoch.
We note that our period determination is in agreement with a value recently
established by \cite{ote08} using the same data from NSVS as used in this paper.
\section{Spectroscopy}
The spectrum of B17 is found to be highly variable on both short ($<$day) and long
($\sim$year) timescales, with the most prominent features in the spectra being
H, \ion{He}{i}, \ion{He}{ii} and \ion{N}{iii} lines in absorption and \ion{He}{ii}
and \ion{N}{iii} lines in emission. Wavelength shifts are observed
for different lines, some of which become double in some spectra suggesting that it
is a binary system with at least one hot component, as revealed
by the presence of \ion{He}{ii} and \ion{N}{iii}.
\subsection{Long Term Variability}
Fig.~2 shows blue-violet spectra (4000-4900\AA) obtained at random phases between 2004 July and 2007 August.
The most prominent elements observed are H, \ion{He}{i},
\ion {He}{ii} and \ion{N}{iii} in absorption and \ion{He}{ii}~$\lambda$4686 and
\ion{N}{iii}~$\lambda$$\lambda$4634-40-42 in emission. The spectra show strong, highly
variable \ion{N}{iii} and \ion{He}{ii} emission lines; in 2004 July 18 these are of
a similar strength but by 2006 June 11 the \ion{He}{ii} line is almost three times
the intensity of the \ion{N}{iii} lines. The \ion{N}{iii} profile is also variable;
it appears as two separate peaks in some of the spectra (2006 June 11,
2006 September 10-11, 2007 July 21) and it is almost completely blended in the
spectra taken on 2004 July 18 and 2006 August 18, although this in part could be due
to the low resolution of the spectrum from 2004 July 18.
The hydrogen lines are also highly variable; this is particularly evident for
H$\gamma$ when compared to the diffuse interstellar band (DIB) at $\lambda$4428.
H$\beta$ shows a P-Cygni profile varying in both width and strength suggesting the
stellar wind is highly variable.
Finally, the \ion{He}{i}~$\lambda$4471/ \ion{He}{ii}~$\lambda$4541 ratio, which is
employed as a standard diagnostic for the spectral type of early stars
\citep{wal90}, is also observed to vary over the course of the
observations.
\begin{figure}[h]
\centering
\label{fig:bluespec}
\resizebox{\columnwidth}{!}{\includegraphics[trim = 0mm 0mm 40mm 0mm, clip]{12123f2.ps}}
\caption{Blue-violet spectra of B17 obtained at different phases over a
three year period, with the most prominent lines labeled.}
\label{blue}
\end{figure}
The most prominent line in the yellow spectra (5600-6000\AA) is the varying P-Cygni
profile in \ion{He}{i}~$\lambda$5875, which blends with the \ion{Na}{i}
interstellar lines (Fig.~3). When compared to the nearby DIBs at 5780-5800$\AA$,
the absorption trough in \ion{He}{i} line was observed to vary by a factor of 4 in
intensity relative to the nearby DIB features, with the emission
component also varying in strength by a factor of 2. Other lines in the yellow
spectra include \ion{C}{iii}~$\lambda$5696 and \ion{C}{iv} around $\lambda$5810.
\begin{figure}[!ht]
\centering
\label{fig:greenspec}
\resizebox{\columnwidth}{!}{\includegraphics[trim = 0mm 0mm 40mm 0mm, clip]{12123f3.ps}}
\caption{B17 yellow spectra showing variability at different phases taken over
a period of three years. Prominent lines are labeled.}
\end{figure}
The most prominent feature in the red spectra (6350-6650\AA) is the H$\alpha$ line
which also demonstrates a highly variable P-Cygni profile which is blended with
\ion{He}{ii} absorption features in all the red spectra, complicating analysis of
the profile. It has a double peak profile which varies in intensities with the redder
peak being between 2 and 4 times stronger than the bluer peak. The depth of the
\ion{He}{ii} line, which presumably contributes to the double peaked morphology
is also variable (Fig.~4). Preliminary calculations for the velocities of the blue
edge of the H$\alpha$ profile ($v_{{\rm edge}}$) have a range of 2260 - 2625 kms$^{-1}$
with an average 2440$\pm$50 kms$^{-1}$. The $v_{{\rm edge}}$ is related to the
terminal velocity of the wind; in the extreme ultraviolet, the terminal velocities
for OB stars are 15\%-20\% smaller than the edge velocities \citep{pri90}. Variations
in the wind profiles of OB stars are thought to be a consequence of highly structured
winds on multiple scales, and while it appears likely that the variations observed
for B17 arise due to a highly anisotropic circumstellar envelope observed at differing
lines of sight throughout the orbital period, the current limited data set offers
little prospect of a more explicit physical interpretation.
\begin{figure}[!ht]
\centering
\label{fig:redspec}
\resizebox{\columnwidth}{!}{\includegraphics[trim = 0mm 0mm 40mm 0mm, clip]{12123f4.ps}}
\caption{B17 red spectra showing H$\alpha$ variability at different phases
taken over a period of three years. Prominent lines are labeled.}
\end{figure}
Unfortunately, no spectra with the same orbital phases have been obtained at
different epochs and so it has not been possible to search for unambiguous long term
secular variability in the current data set.
In Fig. 5 we plot the EW of the \ion{He}{ii}~$\lambda$4686 line against orbital phase, finding the EW to
be broadly anticorrelated with the photometric light curve, increasing by a factor of
two during the entry to eclipse. This result suggests that the strength of the line
remained $\sim$constant over the segments of the orbital period sampled by the
observations, implying that the change in EW is primarily due to dilution by the
variable continuum.
\begin{figure}[!ht]
\centering
\label{fig:EWs}
\resizebox{\columnwidth}{!}{\includegraphics[trim = 0mm 0mm 0mm 0mm, clip]{12123f5.ps}}
\caption{Upper panel: Light curve folded on a 4.0217 day period. Lower panel: Equivalent widths of the \ion{He}{ii}~$\lambda$4686
emission line for the WHT data taken in 2007 August plotted on the same orbital period.}
\end{figure}
\subsection{Short Term Variability}
The set of 12 high resolution spectra obtained during the nights of the 2007 August
21-22 cover the phases $\phi$=0.38-0.453 and $\phi$=0.629-0.701 of one period. The
spectra obtained during the first night (Fig.~6 upper panel) have narrower H and \ion{He}{ii}
absorption lines lines compared to the spectra from the second night. During the first
night, the \ion{He}{i}~$\lambda$4471 /\ion{He}{ii}~$\lambda$4541 ratio - used for
classifying O stars - changes monotonically in the sense of increasing \ion{He}{i}~$\lambda$4471
strength, reaching unity in the final spectrum.
A second, blueshifted absorption component appears on the shoulder of the
\ion{He}{i}~$\lambda$4471 line, becoming stronger with time; similar evolution is also
observed in the H$\delta$ profile. Initially, the \ion{N}{iii}~$\lambda$4634-40-42
emission is highly blended, with the redshifted peak being the strongest. These peaks
start to separate on the last spectrum of that night ($\phi$=0.453). The \ion{He}{ii}~$\lambda$4686
emission feature appears to be non-symmetric and strengthens throughout
the night. The \ion{He}{i}~$\lambda 4026$, \ion{He}{ii}~$\lambda$4200 and H$\gamma$
lines do not vary significantly throughout the first night.
During the 17hrs between the last observation of night one ($\phi$=0.453) and the
first of night two ($\phi$=0.629) the spectrum has clearly evolved. The
\ion{He}{i}~$\lambda$4026, \ion{He}{ii}~$\lambda$4200 and H$\gamma$ lines are broader
and non symmetrical. On the second night (Fig.~6 lower panel),
the \ion{He}{i}~$\lambda$4471 / \ion{He}{ii}~$\lambda$4541 ratio is $>$ 1 on the first spectrum
of the second night and decreases until it reaches unity on the last spectrum
($\phi$=0.701). Significant changes are apparent in the \ion{He}{i}~$\lambda$4471
absorption profile, which is much broader and appears to be composed of two troughs,
with the bluer trough being stronger. An additional, distinct weak blueshifted
absorption line is also present and becomes stronger with time. The H$\delta$ line
appears to be the result of the blending of two lines, showing a wide profile with
two troughs. The bluer trough is stronger than the redder during the second night.
The \ion{N}{iii}~$\lambda$4634-40-42 emission is observed as two separate peaks,
with the redder being the strongest and the intensity of both increasing through
the night. The \ion{He}{ii}~$\lambda$4686 emission feature appears to be symmetric
compared to the first night and weakens throughout the night. We note that (as
mentioned in Section 3) the light curve demonstrates an asymmetry - possibly due to
mass transfer - between phases $\phi \sim$0.6--0.75. Given this is coincident with
the observations during night two, it is tempting to attribute the spectral changes
(in part) to such a process.
\begin{figure*}[h!]
\centering
\label{fig:blueshort}
\includegraphics[trim = 0mm 8mm 0mm 0mm, clip,scale=0.55, angle=270]{12123f6a.ps}
\includegraphics[trim = 0mm 8mm 0mm 0mm, clip,scale=0.55, angle=270]{12123f6b.ps}
\caption{Blue spectra obtained with the WHT on the night of 2007 August 21 and 22.
The main transition lines are shown. On the regions where the lines are blended,
accuracy of the identified lines is not certain and will be clarified with spectra
dissentangling.}
\end{figure*}
\subsection{Spectral Classification}
Given that the current spectroscopic dataset does not span an entire orbital cycle
we have not attempted a formal deconvolution of the blended spectra. Instead we
have simply used the spectra closest to the primary and secondary eclipses to
perform a {\em preliminary} classification of the components of the system,
using the stellar atlas of \cite{wal90} and \cite{wal00}.
The spectrum at $\phi$=0.453 shows a \ion{He}{ii}~$\lambda$4541 /
\ion{He}{i}~$\lambda$4471 absorption-line ratio of one, suggesting a likely spectral
classification of O7. Both \ion{He}{ii}~$\lambda$4686 and \ion{N}{iii}~$\lambda\lambda$4634-40-42
are in strong emission which is characteristic of an
Of supergiant, implying an initial classification of O7~Iaf. Fig.~7 upper panel shows this
spectrum along with comparison spectra. The spectrum lacks the \ion{Si}{iv}~$\lambda$4089
line, which is unexpected for an O7 SG.\footnote{We note that while this line is
not present in the spectrum of the O7 supergiant Sanduleak 80, given the reduced
metalicity appropriate for a SMC star it is not clear it provides a valid comparison.}
\cite{wal01} mentions that a similar effect is found in spectra in the Small Magellanic
Cloud where no \ion{Si}{iv}~$\lambda$4089 is observed and he attributes this to the absorption
and emission features cancelling each other. Therefore we suggest a preliminary spectral
classification of O7~Iaf+ for this component.
\begin{figure}[!ht]
\centering
\label{fig:specclass}
\resizebox{\columnwidth}{!}{\includegraphics[trim = 10mm 9mm 5mm 60mm, clip]{12123f7a.ps}}
\resizebox{\columnwidth}{!}{\includegraphics[trim = 8mm 0mm 5mm 75mm, clip]{12123f7b.ps}}
\caption{Upper panel: B17 spectrum obtained with the WHT on the night of 2007 August 21,
close to the secondary eclipse, compared with Of supergiant spectra from the
Digital Atlas of Stellar Classification \citep{wal90}.
The spectrum is most compatible with an O7Iaf classification. Double lines are observed
even though the system is at eclipse which we currently cannot account for and will be
further studied with the spectra dissentangling. Lower panel: B17 spectrum obtained
with the WHT on the night of 2006 June 11,
close to the primary eclipse, compared with Of supergiant spectra from the Digital
Atlas of Stellar Classification \citep{wal90}. The spectrum is
more comparable to the O9 classification.
}
\end{figure}
The spectrum at $\phi$=0.036 (Fig.~7, lower panel) shows a \ion{He}{ii} $\lambda$4541 /
\ion{He}{i} $\lambda$4471 absorption-line ratio smaller than one, indicating a
spectral type later than O7. The \ion{He}{ii} $\lambda$4686 and \ion{N}{iii}
$\lambda$$\lambda$4634-40-42 are also in strong emission, implying an Of
supergiant. The spectrum also presents a stronger Si \textsc{iv} 4089 absorption
line compared to the spectrum at $\phi$=0.453 although it is still weaker than
expected for a late O star. Therefore we suggest a preliminary spectral
classification of O9~Iaf for this component.
Since the combination of spectra of the preliminary
classifications of an O7 and an O9 components would not give a HeI4471/HeII4541 ratio less
than unity (see Section 4.2), it suggests there is emission in-filling of HeI4471.
\subsection{Radial Velocities}
The high resolution blue spectra show systematic night to night variations in the
H, He and N lines, revealing significant radial velocity shifts. Radial velocities
were determined for the primary lines which appear not to be blended
(\ion{He}{ii}~$\lambda$4200, \ion{He}{ii}~$\lambda$4541, H$\gamma$~$\lambda$4340) in the high resolution
spectra obtained on the WHT observing run in 2007 August. Gaussians were fitted to the line profiles using
the {\sc dipso} emission line fitting command ({\sc elf}; the results are discussed in Sec.~5). The lines for the
secondary appear to be blended and were not measured. The radial velocities for the secondary will be
measured in the future paper after the spectra have been dissentangled.
\section{Light and radial velocity curve analysis}
In order to reproduce the observed characteristics of the photometric light curve, we analysed it with
the 2003 version of the \citet[W-D;][]{wil71} code.
For the analysis, the ROTSE (with an effective wavelength similar to Johnson $R$)
and amateur $V$ light curves were modeled as different datasets. In all cases,
detailed reflection-model and proximity-effect corrections were included.
Considering the spectroscopic analysis, the temperature of the primary ($T^{\rm
P}_{\rm eff}$) was fixed to 35\,000~K, and the bolometric albedo and gravity
brightening coefficients were set to unity, as generally found for stars with
radiative envelopes. In addition, a circular orbit was adopted, as suggested by
the equal separation between both (primary and secondary) eclipses, and a
rotation rate synchronized with the orbital period was assumed for both
components. The fitting process was carried out iteratively until three
consecutive solutions provided differential corrections for all the parameters
smaller than twice their internal errors.
Considering the characteristics in the light curves described above, the first
runs in the modelling with W-D assumed a {\em semi-detached} configuration.
Numerous attempts were performed with several mass ratios and with either the
primary or the secondary component filling the Roche lobe. However, {\em in all cases,
the fits provided solutions where both stars tended to be in contact}.
Therefore, despite the properties of the lightcurve suggesting a semi-contact configuration
we finally attempted an over-contact solution - the results of which are described
here - where each component could have a different temperature, as observed
from the different depths of the eclipses. For this model, the mass ratio was
set to $q=0.75$ - as might be expected for O7 and O9 super-giant components - noting that
the mass ratio cannot be smaller than ~0.4\footnote{For a mass ratio of 0.4, the implied mass of the primary
-considering the minimum amplitude of the radial velocity curve- is over 120 M$_{\odot}$, regardless of the
light curve analysis.}. The time of minimum ($t_{\rm min}$), the period ($P$), the orbital inclination
($i$), the temperature of the secondary component ($T^{\rm S}_{\rm eff}$), the
surface potential ($\Omega^{\rm P}$) and the luminosity of the primary
component ($L^{\rm P}$) were all left as free parameters in the fit.
The $V$ light curve (which has smaller photometric errors) clearly reveals that one quadrature is
$\sim$0.02 mag brighter than the other. Two main solutions were attempted to
explain the observed O'Connell effect, one with an equatorial hot spot (30\%
hotter than the photosphere) on the primary component and another one with the
spot being on the secondary component. The size and position of the spot were
left as free parameters and also fitted at each run, although we emphasise that the
size and temperature of the spot are strongly correlated. The fits with the spot on
the primary component were finally adopted, since they provided slightly
smaller errors. In addition, the position of the spot (oriented roughly towards
the secondary component) can more easily be explained as the interaction of the
stellar winds.
Together with the light curve analysis, the radial velocities were also used to
constrain the solution. Radial
velocities were fitted separately from the light curve to avoid the larger
number of photometric observations dominate in the solution; the parameters fitted
being the semi-major axis ($a$) and the systemic velocity
($\gamma$). The rms of the fit is 9.5 km~s$^{-1}$ for the radial
velocities, 0.023 mag for the $V$ light curve and 0.048 mag for the ROTSE light
curve. Unfortunately, the accuracy of the solution obtained is strongly dependent on the
mass ratio adopted. Nevertheless, Fig.~8 and Fig.~9
show the light curve and radial velocity curve with the best fit and Table \ref{lcparam}
gives the parameters obtained from this solution.
Despite our efforts we are still not completely satisfied with this fit. In particular, we
are still unable to adequately fit the egress from secondary minimum nor the
subsequent lightcurve between phases 0.6-0.9. Moreover it is expected that the temperatures
of stars in over-contact binaries should be $\sim$equal where a ratio of 0.85 was found between
the secondary and primary. As such we regard the parameters presented in Table 2 as provisional
at present. We anticipate that the determination of radial velocities for the secondary
component, which will directly constrain the binary mass ratio
will greatly clarify the fundamental properties of the components and
the configuration of this EB system.
Finally, the lightcurve modelling, allows us to address the distance to B17, and by extention
the Cyg OB2 association. Adopting the bolometric
corrections from \cite{mar05} we may use the bolometric luminosities determined above (See Table 2) to
calculate the absolute $V$ magnitudes of both components.
The $V$ band reddening was then calculated by following
a similar procedure used by \citealt{neg08} (see Section~6)
With the absolute $V$ magnitudes, and the $V$ band reddening for B17, along with the $V$ band values for
the two minima in the lightcurve, the distance modulus was calculated to be 10.9-11.3 which
corresponds to a distance of about 1.5-1.8~kpc. While we regard these values as provisional due to
the difficulties in modelling described above, these are consistent with the
commonly adopted distance estimate of 1.7~kpc. (e.g. \citealt{han03}, \citealt{tor91} and \citealt{kim07}). Therefore they do
not agree with the distance estimate of 900-950~pc obtained by \cite{lin09} from
the lightcurve modelling of Cyg OB2 $\#$5.
A distance of 900~pc would imply radii that are $\sim$half those currently derived from the modelling.
The derived radii scale linearly with the semi-major axis and the semi-major axis in turn is strongly
dependant on the assumed mass ratio. For this system, a semi-major axis a factor of 2 smaller would
imply a mass ratio of 6. We consider this to be extremely unlikely, but not impossible.
\begin{table}
\caption {Results from the analysis of the light and radial velocity curves: The errors shown
should be considered internal errors of the fit. Any possible systematic errors
are not included.}
\label{lcparam}
\centering
\begin{tabular}{l c c }
\hline\hline
Parameter & Value\\
\hline
T$_{0}$ (MJD) & 4272.534 $\pm$ 0.004
\\
Period, $P$ & 4.02174 $\pm$ 0.00003 days
\\
Inclination, $i$ & 72 $\pm$ 1.5
\\
Eccentricity, $e$ & 0 (Fixed)
\\
Mass ratio $q$ & 0.75 (Fixed)
\\
Temperature ratio & 0.85 $\pm$ 0.02
\\
Semi-major axis ($R_{\odot}$) & 50 $\pm$ 1
\\
Systemic velocity & -43 $\pm$ 4 km s$^{-1}$
\\
\hline
\\
\hline\hline
Parameter & Primary & Secondary\\
\hline
Mass ($M_{\odot}$) & 60 $\pm$ 5 & 45 $\pm$ 4
\\
Radius ($R_{\odot}$) & 22 $\pm$ 1 & 19 $\pm$ 1
\\
log g (cgs) & 3.53 $\pm$ 0.01 & 3.52 $\pm$ 0.01
\\
RV Semi-amplitude (km s$^{-1}$) & 257 $\pm$ 7 & 343 $\pm$ 9
\\
mean Teff (K) & 35000 (Fixed) & 29900 $\pm$ 700
\\
Surface potential ($\Omega$) & 3.14 $\pm$ 0.02 & 3.14 (Fixed)
\\
M$_{Bol}$ & -9.8 $\pm$ 1 & -8.8 $\pm$ 1
\\
\hline
\end{tabular}
\end{table}
\begin{figure}[h!]
\centering
\label{fig:lcfit}
\resizebox{\columnwidth}{!}{\includegraphics{12123f8.ps}}
\caption{The $V$ band and $R$ band light curves of B17 with the best fit model overlaid.
The panels below the light curves show the deviations from the model.}
\end{figure}
\begin{figure}[!ht]
\centering
\label{fig:rvfit}
\resizebox{\columnwidth}{!}{\includegraphics[trim = 0mm 70mm 0mm 0mm, clip]{12123f9.ps}}
\caption{The RV curve for the primary component obtained from the WHT data, with the best
fit overlaid (solid line). The panel below shows the
residuals between the observed data and the model.}
\end{figure}
\section {Discussion}
Given the difficulties in determining a unique model fit to the data (Sect. 5), we regard
the parameters presented in Table~2 to be provisional. As a check on consistency we
have calculated a preliminary bolometric magnitude for the components, following a similar procedure
the that employed by \cite{neg08}. Assuming that both
components are of similar colour\footnote{The colour differences between an O7
and an O9 supergiant are $(J-V)_{0}=0.09$ and $(K-V)_{0}=0.11$ \citep{weg94}},
we adopted the 2MASS value for $(J-K_{{\rm S}})$ used in \cite{neg08},
the effective temperature $T_{{\rm eff}}$ and bolometric correction (BC)
calibrations of \cite{mar05}, and the intrinsic $(J-V)_{0}$ and
$(K-V)_{0}$ colour calibrations of \cite{weg94}. Using the 2MASS observed
$(J-K_{{\rm S}})$, we derived $E(J-K_S)$. The reddening of the system was
calculated using the relation $A_{K_{{\rm S}}}=0.67E(J-K_{{\rm S}})$, due to the
reddening in the association being close to standard \citep{han03}. The reddening
in the $V$ band was calculated using the relation $0.112A_{V}$$\simeq$$A_{K}$
\citep{rie85}. The absolute $V$ band magnitude for the primary was then
calculated, using the $V$ band value of the secondary minimum and adopting the
distance modulus of 11.3, obtained by averaging spectroscopic distances
\citep{kim07}. A semi-observational $M_{{\rm bol}}$
was then calculated to be
$-9.8\pm0.2$ and $-9.2\pm0.2$ for the primary and secondary, by adding $(V-K)_{0}$ and
the BC to the $V_0$, assuming an uncertainty of one spectral type. With this value, the
luminosities were calculated to be log($L_{1}/L_{\odot})=5.8\pm0.1$ and
log($L_{2}/L_{\odot})=5.6\pm0.1$, noting that the main
source of error is likely to be the uncertainty in the spectral type and hence
temperature and BC. We regard these estimates as upper limits since it assumes
that both components are completely eclipsed during the minima.
The positions of the two components of B17 in the HR diagram (Fig.~10) are consistent with other known members of
Cyg OB2 suggesting that it too is a {\em bona fide} member, with an age of $\sim$2.5Myr \citep{neg08}.
With a likely spectral type of O9~Iaf, the secondary appears slightly more
evolved than the primary and hence was likely the initially more massive star,
with both evolving from very early O main sequence stars. We note that a system
composed of O7~Iaf \& O9~Iaf stars is also compatible with the other massive
evolved binaries in Cyg OB2. Indeed, the 6.6~day Ofpe/WNL+O6.5-7 binary Cyg OB2 $\#$5
\citep{lin09} appears to be remarkably similar to B17, although with the Ofpe/WNL star
being slightly more evolved than the O9~Iaf
secondary in B17. The slightly longer period of Cyg OB2 $\#$5 can be explained by one or more from
(i) initial birth parameters, (ii) mass transfer from one star to the other and
(iii) mass lost by both stars via stellar winds.
\cite{mar07,mar08} used observations of the Galactic Centre and Arches
clusters to examine the evolutionary pathways of stars at galactic metallicity with masses in excess of
30M$_{\odot}$, suggesting the following evolutionary pathways \citep[also see][]{cro95}:\\
$\bullet$ { $\sim$30-60M$_{\odot}$ : O $\rightarrow$ Ofpe/WN9 $\rightleftharpoons$ LBV $\rightarrow$ WN8 $\rightarrow$ WN/C}
\\
$\bullet$ { $\sim$60-120M$_{\odot}$ : O $\rightarrow$ Of $\rightarrow$ WNL + abs $\rightarrow$WN7}
\\
With an age of $\sim$2.5 Myr and spectral classifications of O7 and O9, B17 appears to lie at the dividing point between the 2 evolutionary pathways.
However, given that it
appears significantly less massive than the WN6ha + WN6ha binary WR20a
(m$_1$=m$_2$=83$M_{\odot}$ \citealt{rau04,bon04}) we suspect that
it will instead evolve to resemble Cyg OB2 $\#$5 and hence to a configuration similar
to GCIRS 16SW, composed of two cooler extreme B supergiants/LBV candidates
($P_{{\rm orb}}=19.45$ days; \citealt{mar06}),
with stellar mass loss resulting in an eventual lengthening of the orbital period
of B17. Indeed, assuming the system avoids merger during the LBV phase and,
remaining bound, receives a favourable SNe kick to reduce the orbital separation
it might briefly form a high mass X-ray binary with a WR mass donor prior to the second SN.
Irrespective of its ultimate fate, B17 is amongst the brightest/most massive
systems in Cyg OB2 and adds to the increasing number of massive binaries
identified within it (\citealt[][see Table 3]{kim09}). Similar trends for both a high binary fraction
and the multiplicity of the brightest/most evolved cluster members
have been observed for Pismis 24 \citep{mai08}, NGC3603 \citep{sch08},
Westerlund 1 (\citealt{cla08}; \citealt{rit09}) and potentially the Arches \citep{cla09}.
If these trends continue
it will have significant implications for the formation channels and
relative production rates of both low and high mass X-ray binaries and systems
comprising of two relativistic objects \citep{kob07}.
\begin{figure}[!ht]
\centering
\label{fig:cygob2}
\resizebox{\columnwidth}{!}{\includegraphics[trim = 0mm 0mm 0mm 0mm, clip]{12123f10.ps}}
\caption{Updated semi-observational HR diagram from \cite{neg08}, based on
published spectral classes and 2MASS $JHK_{{\rm S}}$ photometry and a DM of 11.3. The continuous lines
are non-rotating isochrones for $\log t=6.2$, $6.3$ and $6.4$ from \cite{sch92} and
the dashed line is the $\log t=6.4$ isochrone in the high-rotation models from
\cite{mey03}. The circle and the triangle show the positions of the primary and secondary components of B17.}
\end{figure}
\begin{table*}
\caption {Evolved massive binaries in Cyg OB2 from Kiminki et al (2009).}
\label{table:kiminki}
\centering
\begin{tabular}{l l l l}
\hline\hline
Star & Sp Types & Period (days) & References \\
\hline
MT05 & O9\,III \& \it{mid B} & 25.1399 (0.0008) & \cite{kim09} \\
MT720 & \it{early B \& early B} & $<$5 & \cite{kim09} \\
Schulte 3 & O6\,IV \&O9\,III & 4.7464 (0.0002) & \cite{kim09}, Kinemuchi et al. in prep\\
Schulte 5 ($\#$5) & O7\,I \& Ofpe/WN9 & 6.6 & \cite{wil48}, \cite{wil51}, \cite{mic53} \\
& & & \cite{wal73}, \cite{con97}, \cite{rau99} \\
Schulte 8a ($\#$8a) & O6\,If \& O5.5\,III(f) & 21.9 & \cite{rom69}, \cite{deb04} \\
Schulte 9 ($\#$9) & O5.5\,If \& O6-7 & 2.35 yrs & \cite{naz08} \\
Schulte 73 & O8\,II \& \it{O8?} & 17.4 (0.2) & \cite{kim09} \\
B17 & O7\,Ia \& O9\,I & 4.0217 (0.0004) & This paper \\
\hline
\end{tabular}
\end{table*}
\section{Summary}
Using photometric and spectroscopic data, we have demonstrated that B17 is an
eclipsing, double lined spectroscopic binary comprising two supergiants with
preliminary classifications of O7Iaf and O9Iaf. The spectra are highly variable, and with a subset revealing
features from both stars, raise the possibility of achieving a dynamical mass
determination for both components. Utilising both the photometric lighturve and our
limited RV dataset we attempted to determine an initial orbital solution for the binary.
Despite the
morphology of the lightcurve indicating a semi-contact configuration we were unable to
to achieve convergence for such a hypothesis and hence were forced to adopted an over-contact configuration.
In the absence of a full RV curve for both system components we were forced to fix the binary mass ratio, and had to
include the presence of a star spot to address the observed asymmetries in the lightcurve (which are likely due to the effects of
binary mass transfer). However, we were
still unable to fully fit both secondary eclipse and the lightcurve between orbital phase $\sim$0.6-0.9
with such a model; as such we regard the modelling results presented in this work as provisional. We anticipate
that a full RV curve for both components will be necessary to obtain more precise parameters for the system;
additional data to accomplish this goal are currently being obtained and refined analysis will be
presented in a future paper.
Nevertheless, the provisional distance calculation of 1.5-1.8 kpc obtained from the
lightcurve analysis agrees with previously published values for the distance
to Cyg OB2, being inconsistent with the distance of 900-950 pc
determined by \cite{lin09} with the Cyg~OB2 \#5 light curve analysis.
When placed in the HR diagram, B17 appears to be consistent with the age and
stellar population of the Cyg OB2 association. Assuming the system avoids
merger, it is likely to evolve through an extreme B
supergiant/LBV phase into a long period WR+WR binary configuration
as mass loss via stellar winds increases the orbital separation.
In combination with the recent work of \cite{kim09} and \cite{kob07} the results of our analysis
provides additional evidence that Cyg~OB2 has a very high fraction of massive binary stars. Such an observational constraint
needs to be considered when determining the initial mass function of the association as it may both influence the slope
of the relationship and also lead to a population of artificially
massive stars, resulting in the inflation of a putative high mass cut-off to the IMF.
\begin{acknowledgements}
We thank Pedro Pastor and Manuel M\'{e}ndez for having obtained and reduced the photometric data. Amparo Marco for help with the 2004 observing run and Miriam Garc\'{i}a for assistance with the 2006 INT run. We thank Dan Kiminki, Fraser Lewis and Chris Evans for useful discussions and reading of the manuscript. We also thank the referee for his guidance in completing the paper for publication. The Faulkes Telescope Project is an educational and research arm of the Las
Cumbres Observatory Global Telescope Network (LCOGT). VS acknowledges support
from the Dill Faulkes Educational Trust.
This research is partially supported by the Spanish Ministerio de Ciencia e
Innovaci\'on undergrants AYA2008-06166-C03-03 and Consolider-GTC CSD2006-70.
The G.D. Cassini telescope is operated at the Loiano Observatory by the
Osservatorio Astronomico di Bologna. The WHT is operated on the island of
La Palma by the Isaac Newton Group in the Spanish Observatorio del Roque
de Los Muchachos of the Instituto de Astrof\'{\i}sica de Canarias. The 2006
observations were taken as part of the service programme
(programme SW2005A20).
\end{acknowledgements}
|
\section{Introduction}
Diffuse interstellar bands are absorption features observed in
starlight crossing diffuse interstellar clouds. Since their discovery
in the beginning of the 20th century, scientists have been puzzled
by the origin of these bands that appear both as relatively narrow
and rather broad bands covering the UV/VIS and NIR
\citep{Tielens:1995}. In the last decennia the idea has been
established that it is unlikely that all these bands are due to one
or a very few carriers and with the progress of optical laboratory
techniques, several families of potential carriers have been
investigated.
It was shown that the electronic transitions of a series
of PAH-cations do not match the listed DIBs
\citep{Salama:1996,Salama:1999,Brechignac:1999,Ruiterkamp:2002}.
Similarly, systematic laboratory studies of electronic spectra
of carbon chain radicals have not resulted in positive
identifications either
\citep{Motylewski:2000, Ball:2000a,Jochnowitz:2008},
even though it is known from combined radio-astronomical and
Fourier Transform Microwave (FTMW) studies that many of such
species are present in dense clouds \citep{Thaddeus:2001}. Only
C$_3$ has been recorded unambiguously in diffuse interstellar clouds
\citep{Maier:2001}.
Other studies, focusing on multi-photon excitation in molecular
hydrogen \citep{Sorokin:1998}, or spectra of fullerenes and
nano-tubes \citep{Kroto:1992,Foing:1994} have been unsuccessful
as well.
In the past years, several coincidences between laboratory and
astronomical DIB studies have been reported in the literature.
These have all turned out to be accidental, and from a statistical
point of view, the chance of an overlap is also quite substantial,
DIBs cover a major part of the wavelength region between
roughly 350 and 1000 nm. However, there are a number
of conditions that have to be fulfilled before a coincidence
of a laboratory and an astronomical DIB spectrum may be interpreted
as a real match. These conditions have become more and more strict
with the recent improvement in achievable spectral resolution, both
in laboratory and astronomical studies.
\begin{figure}
\centering
\includegraphics[width=9cm]{Figure1.eps}
\caption{The $\lambda$5450 DIB. The top spectrum is a simulated
spectrum as available from DIB catalogues. The middle and bottom
figure show observational spectra of the HERMES and McKellar
spectrograph, respectively.
The HERMES spectra show the $\lambda$5450 DIB recorded
toward HD 183143 and toward a reference star (HD 164353). The
McKellar spectra show a reference spectrum toward Rigel (top),
the DIB spectrum also toward HD 183143 (bottom) and the corresponding
spectrum (middle) in which the SII stellar line has been deblended.
}
\label{observational spectra}
\end{figure}
The two most important DIB matching criteria to link laboratory and
astronomical data are:\\
\begin{enumerate}
\item The gas phase laboratory and observational values of both peak
position and bandwidth of the origin band transition should be
identical, unless it can be argued that a spectral shift or band
profile change may be due to an isotope or temperature effect.
An example of the latter is given by spectroscopic measurements
on benzene plasma yielding an absorption feature coinciding with the
strongest DIB at 442.9 nm \citep{Ball:2000b,Araki:2004}.
The laboratory FWHM turned out to be narrower than in the astronomical
spectrum. It was argued that the spectrum of a non-polar molecule
cooled in a molecular expansion may be considerably colder
than in space where only radiative cooling applies. A similar
discussion has been given by \citep{Motylewski:2000} who showed that
unresolved rotational profiles may change substantially for different
temperatures, as has also been calculated and discussed by
\cite{Cossart:1990}.
\\
\item Once the origin band overlaps with a DIB feature, gas phase
transitions to vibrationally excited levels in the electronically
excited state of the same carrier molecule should match as well and
the resulting band profiles should behave in a similar way (i.e. with
comparable equivalent width ratios) \citep{Motylewski:2000}.
A good example for this is the electronic spectrum of C$_7$$^-$ that
has been regarded for several years as a potential carrier as
subsequent electronic bands fulfilled both conditions
\citep{Tulej:1998,Kirkwood:1998}. Detailed follow-up studies showed
that the series of (near) matches was coincidental \citep{McCall:2001}. \\
\end{enumerate}
In the end, and despite much progress both from the observational and
laboratory side, all efforts to assign DIBs have resulted in a rather
static situation -- triggering more and more exotic explanations for
DIB carriers -- and the origin of the DIBs is still as mysterious as it
was nearly 100 years ago.
In this letter we report a match of a laboratory spectrum with a
diffuse interstellar band that is special as the first condition
is fulfilled for a rather broad and potentially lifetime broadened
DIB, i.e. the laboratory and astronomical spectrum should be fully
identical, independent of temperature restrictions. New astronomical
observations obtained with the Mercator telescope, using
the HERMES spectrograph and the Dominion Astrophysical Observatory
(DOA) 1.2 meter telescope, using the McKellar spectrograph are
presented in order to characterize the band profile of the
$\lambda$5450 DIB with the best possible resolution.
Even though we have not been able to unambiguously identify
the laboratory carrier, that most likely is a smaller hydro-carbon
bearing molecular transient, we think that this overlap is important
to report, as it provides a new piece of the puzzle.
\section{Laboratory Experiments}
The experimental set-up has been described
\citep{Linnartz:1998, Motylewski:1999} and has been extensively used
to study a large number of carbon chain radicals of astrophysical
interest \citep{Jochnowitz:2008}. The monochromatic output $\sim$
0.1\ cm$^{-1}$ at 540 nm ($\sim$ 18,500\ cm$^{-1}$) of a pulsed dye
laser based cavity ring-down setup is focused into an optical cavity
consisting of two highly reflective mirrors (R $>$ 0.9999). A
special pulsed high pressure slit-nozzle system capable of producing
intense 300 $\mu$s long plasma pulses by discharging (- 1 kV, 100 mA)
an expanding gas mixture of 1 $\%$ acetylene (C$_2$H$_2$) in He
is mounted inside the cavity with its slit parallel to the optical
axis of the cavity. In the expansion a large variety of new species
is formed and as the technique is not mass selective, special care
has to be taken when assigning bands to specific carriers. Mass
selective matrix isolation spectra offer a good starting point for
an assignment \citep{Jochnowitz:2008}. In the case of rotationally
resolved spectra unambiguous identifications are generally possible,
either by combination differences of accurate spectral fits, or by
isotopic studies using C$_2$D$_2$ instead of C$_2$H$_2$ (or a mixture
of C$_2$H$_2$/C$_2$D$_2$). The source runs at 30 Hz and special care is
taken that the pressure inside the cavity remains constant during jet
operation to reduce baseline fluctuations. Rotational temperatures
are typically of the order of T$_{rot}$ $\sim$ 10\--20 K. This low
temperature results in a spectral simplification and simultaneously
increases the detection sensitivity because of an improved
state-density. In addition, the source offers a Doppler free
environment with a relatively long effective absorption path length.
The laser beam intersects the 3 cm long planar expansion about
5\--10 mm downstream using a sophisticated trigger scheme. Subsequent
ring-down events (typically 20-30 $\mu$s for a 52 cm long cavity)
are recorded as function of the laser frequency by a photo-diode
and transferred to an averaged ring-down time by fitting 45
subsequent ring-down events. This value as function of the
laser wavelength provides a sensitive way to record optical spectra.
An absolute frequency calibration is obtained by recording an I$_2$
reference spectrum simultaneously.
\section{Astronomical Observations}
The laboratory data are compared to observations from two different
astronomical facilities.
\subsection{HERMES @ Mercator Telescope}
The HERMES observations were carried out in service mode using the
Mercator telescope at Roque de los Muchachos Observatory on La Palma.
The 1.2 m telescope is operated by the Katholieke Universiteit in
Leuven, Belgium, in collaboration with the Observatory in Geneva,
Switzerland.
The spectra were obtained in June 2009 with HERMES (High Efficiency
and Resolution Mercator Echelle Spectrograph) \citep{Raskin:2008},
which is a fibre-fed-cross-dispersed spectrograph.
The spectrograph has a fixed spectral format and samples the spectrum
between 377 and 990 nm in 55 spectral orders on a 4.6\ k x 2\ k CCD.
The spectral resolution is slightly variable over the field,
but is 85,000 on average. We obtained 3 spectra of 1,200\ s of
HD 183143 (B7Ia, m(v)=6.92, B-V=+1.001), the DIB spectral standard
with a reddening E(B-V) close to 1.0. The reference star HD 164353
(B5Ib, m(v)=3.97, B-V = $-$0.002) was sampled in 3 exposures of 1 min.
The spectral reduction was performed using the specifically coded
HERMES pipeline and contains all standard steps in spectral reduction.
The wavelength calibration is based on spectra of ThAr and Ne lamps.
As we are mostly interested in the broad absorption feature that
is centred around 545 nm, we focus further on this spectral region
of HD 183143. The spectra are shown in Figure 1 (middle rows) and
compared to the $\lambda$5450 DIB profile as available from a
series of digital DIB catalogues
\citep{Herbig:1975, Jenniskens:1994, Tuairisg:2000, Galazutdinov:2000}
in the upper row.
\subsection{McKellar @ DAO Telescope}
Fifty-five half-hour spectra were taken with the McKellar Spectrograph
and SITe-4 CCD at the DAO 1.2 m telescope, operated by the National
Research Council of Canada, over 6 nights between 16 and 23 July 2006
(UT) at a dispersion of 10.1 $\AA$/mm giving 0.151 $\AA$/pixel
for a resolution $\sim$ 0.3 $\AA$. The data were processed in a
standard fashion using IRAF
\footnote{IRAF is distributed by the National
Optical Astronomy Observatory, which is operated by the Association of
Universities for Research in Astronomy (AURA) under cooperative agreement
with the National Science Foundation.}.
The aggregate spectrum had a
signal to noise of about 1200/pixel before correction of telluric lines.
Removal of the quite weak telluric features was performed conventionally
with spectra (S/N $\sim$1600) of the A0 V star zeta Aql (HD
177724) as the template.
Rigel, an unreddened comparison star with a B8 Ia spectral type very
similar to the B7 Ia of HD 183143, was also observed in order to identify
photospheric lines which contaminate the interstellar features observed in
the latter star. The sharp line at approximately 5454\AA\ arises from S
II and was removed from the spectrum of HD 183143 by simply fitting a
Voigt profile to the line and subtracting this from the original spectrum.
The final ``deblended'' spectrum is plotted as a comparison in Fig. 1
(lower panel, middle spectrum).
\section{Results}
In Fig. 2 several spectra in the 543-547 nm region are compared.
The top spectrum is the digital DIB spectrum of the $\lambda$5450
DIB \citep{Herbig:1975, Jenniskens:1994, Tuairisg:2000, Galazutdinov:2000}.
The spectrum in the middle is a zoom in on the
deblended McKellar spectrum as shown in Fig. 1. The bottom spectrum
is the laboratory spectrum recorded in direct absorption through
an expanding 1$\%$ C$_2$H$_2$/He plasma. The similarity between
the three spectra is striking.
This wavelength region was initially scanned to search for the
$^1$$\Pi$$_u$\ --\ X$^1$$\Sigma$$_g$$^+$ electronic origin band
spectrum of the linear carbon chain radical C$_7$ (following the
C$_7^-$ DIB discussion) that was located in matrix isolation
experiments around 542.3 nm. The laboratory spectrum, shown in
Fig. 2, consists of many narrow lines that are due to small
acetylene fragments (typically C$_2$ and CH) that get weaker
when the distance from the nozzle orifice to the optical axis is increased,
but there is clearly a broad feature lying underneath.
As this band shifts by 1.5 nm to the red upon C$_2$D$_2$ precursor
substitution, it initially was neglected, as for C$_7$ both
C$_2$H$_2$ and C$_2$D$_2$ should result in an identical spectrum.
The shift is illustrated in Fig. 3. In addition, the deuterated spectrum
appears to be somewhat stronger. Despite this
negative result for C$_7$ the profile hiding under the narrow
lines in the C$_2$H$_2$ precursor experiment perfectly matches the
$\lambda$5450 DIB as available from the DIB databases, reason
why additional observations were performed.
\begin{figure}
\centering
\includegraphics[width=9cm]{Figure2.eps}
\caption{The top spectrum shows the digital $\lambda$5450 DIB, the
middle spectrum shows the deblended McKellar data and
the bottom spectrum shows the laboratory cavity ring-down absorption
spectrum through a supersonically expanding acetylene plasma.}
\label{}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=9cm]{Figure3.eps}
\caption{Comparison between laboratory experiments sampling
expanding plasma using regular acetylene (top) and deuterated
acetylene (bottom) as a precursor gas.}
\label{observational spectrum and simulation}
\end{figure}
\section{Discussion}
There is little discussion possible about the coincidence between the
recorded laboratory spectrum and the $\lambda$5450 DIB. Both bands
have a central peak position of 545 nm and a FWHM of 1.03 (0.1) nm
(laboratory spectrum) and 0.953 nm (observational spectrum)
\citep{Tuairisg:2000}. The uncertainty in the first value is due to
the overlap of the many individual transitions that prohibits a clear
view on the broad feature. The question is more whether this actually
represents a DIB match and for this complementary information is
needed. Additional laboratory work has been performed, where it
should be noted that scans as shown in Figs. 2 and 3 typically last
45 minutes to an hour, in order to achieve the required sensitivity
and to cover a frequency domain large enough to discriminate band
profile and base line, i.e. fast optimizations are not possible.
The laboratory band does not show any structure that can be related
to unresolved P, Q and R-branches. With 1.03 (0.1) nm the band is
also much broader than the unresolved rotational profile of a larger
carbon chain radical. For comparison, at 15 K, the band profile of
the linear C$_6$H radical (at 525 nm) is about five times narrower
\citep{Linnartz:1999}. It should also be noted that such a broad
feature actually represents a large absorption compared to many of
the sharper DIBs. Changing the experimental settings to vary
the final temperature in the expansion by measuring close
($\sim$ 50\ K\ $^{\prime}$warm$^{\prime}$) and far
($\sim$ 10\ K\ $^{\prime}$cold$^{\prime}$) downstream, does not
substantially change the FWHM of the spectral contour.
As the narrow overlapping transitions have FWHMs close to the
laser bandwidth, experimental broadening artifacts such as residual
Doppler broadening in the expansion or amplified spontaneous emission,
can be excluded. It is clear that the band profile is due to a
temperature independent and carrier specific broadening effect,
presumably life time broadening. The observed bandwidth of 1.0 nm
($\sim$ 35\ cm$^{-1}$ around 545 nm) corresponds to a lifetime
of roughly 0.15 ps.
The bandwidth profile does not allow concluding on the nature of
the laboratory carrier. The carrier must be a transient species
(a molecular radical, a cation or anion, a weakly bound
radical complex, possibly charged, or a vibrationally or
electronically excited species) as no comparable spectra are
recorded without plasma (i.e. with a regular C$_2$H$_2$/He
expansion). The use of a C$_2$D$_2$/He expansion results in
a red--shifted spectrum (Fig. 3) and from this it can be concluded
that the laboratory carrier must contain both carbon and hydrogen.
In order to check whether there are equivalent H-atoms in this
carrier a C$_2$H$_2$/C$_2$D$_2$ 1:1 mixture in He has been used
as an expansion gas, but this only results in a very broad
absorption feature covering the whole region between results
obtained from pure C$_2$H$_2$ and pure C$_2$D$_2$ expansions.
It is not possible, as demonstrated for HC$_6$H$^+$ or HC$_7$H
\citep{Sinclair:1999, Ball:2000a, Khoroshev:2004}, to conclude
on the actual number of equivalent H-atoms in the carrier by
determining the number of bands that shows up. Also the use
of another precursor (e.g. allene) did not provide conclusive information.
Additional experiments have been performed. The 543-545 nm region
has been scanned using a two-photon REMPI-TOF experiment with the
aim to determine the mass of the carrier \citep{Pino:2001}. No
spectrum could be recorded, which may be related to the short
lifetime of the excited state or with the fact that the carrier
is an ion. Ions are indeed formed in this planar plasma source
\citep{Witkowicz:2004}. Both smaller and larger species have
been observed, with optimum production rates depending, among
other things, on the backing pressure. The production of larger species is
generally more critical, e.g. higher backing pressures are needed
but this also may destabilize the plasma which is unfortunate,
particularly during long scan procedures. More complex species
are generally found further downstream, but in this specific case
we did not observe large differences as function of the distance
from the laser beam to the nozzle orifice. This is the typical
behaviour for a smaller constituent in the gas expansion. We have
tried to study systematically the voltage dependence of the signal;
for a positive ion an increase in voltage should go along with a
decrease in signal for distances further downstream, as the jaws
carry a negative voltage. For anions it is the opposite, but 10 years
of experience with this source have shown that negative ions are rather
hard to produce. Again, the changes we recorded were small and did
not allow drawing hard conclusions. Following condition 2 mentioned
in the introduction, we have also searched in other wavelength regions
blue shifted by values typical for an excited C--C, C$=$C, C$\equiv$C
or CH stretch in the upper electronic state. Such excited bands
have not been observed here, but it should be noted that these
bands can be intrinsically weak. \\
In summary, we are left with a laboratory spectrum that coincides
both in band maximum and band width with a known DIB band at 545 nm.
Our measurements show that the absorption spectrum of a transient
molecule containing hydrogen and carbon reproduces the astronomical
spectrum. The profile can be explained with life time broadening
and this is consistent with the observation that the laboratory
and astronomical spectrum are identical, i.e. without temperature
constraints. In addition, it explains why the large bandwidth of
this DIB does not vary along different lines of sight. The large
effective absorption also may be indicative for an abundant carrier.
The exact carrier, as such, remains an open question.
The present result, however, may be useful to stimulate upcoming DIB work.
\begin{acknowledgements}
The results presented here bridge a period of 10 years. The cavity
ring-down measurements were performed in the Institute for Physical
Chemistry (Department of Chemistry, University of Basel) with
support of the Swiss National Science Foundation and the analysis
follows recent observations and a collaboration within the framework
of the FP6 research training network {\it The Molecular Universe}.
Additional financial support of NOVA is gratefully acknowledged.
\end{acknowledgements}
|
\section{Introduction}
The $B$ physics programme continues to play a crucial role in testing the CKM mechanism of quark flavour mixing and in determining fundamental Standard Model parameters. In recent years accurate measurements of numerous $B$ physics observables have been realized, which calls for an equal improvement in refining the theoretical expectations within and beyond the Standard Model.
The main obstacle for precise theoretical predictions are the complicated strong-interaction effects encoded in the hadronic matrix elements. The QCD dynamics often simplifies considerably in the heavy quark limit $m_b \gg \Lambda_{QCD}$, which allows to establish factorization theorems that disentangle short- and long-distance effects. This separation provides the key for a systematic improvement of the theo\-retical predictions by computing higher order radiative corrections, which should be supplemented by similar progress in the determination of the remnant non-perturbative hadronic parameters.
Here we report on recent progress in the perturbative calculations for charmless hadronic and semileptonic $B$ meson decays. As a technical account of these calculations has already been presented in~\cite{Bell:2009rg}, we focus here on the phenomenological implications of the NNLO corrections.
\section{Hadronic B decays}
Most of the observables at current and future $B$ physics experiments are related to hadronic two-body decays. Among these time-dependent and direct CP-asymmetries in penguin-dominated decay modes are of particular phenomenological interest due to their sensitivity to New Physics.
The control of the strong-interaction dynamics in non-leptonic decays is obviously demanding. The factorization formula for the hadronic matrix elements of the operators in the weak effective Hamiltonian takes a twofold structure~\cite{QCDF},
\begin{eqnarray}
\langle M_1 M_2 | Q_i | \bar{B} \rangle
&\;\simeq\; &
F^{B M_1}(0) \; f_{M_2}
\int du \;T_{i}^I(u) \; \phi_{M_2}(u)
\nonumber \\
&&
+ \; \hat{f}_{B} \; f_{M_1} \; f_{M_2}
\int d\omega dv du \; T_{i}^{II}(\omega,v,u)
\; \phi_B(\omega) \; \phi_{M_1}(v) \; \phi_{M_2}(u),
\end{eqnarray}
which consists of universal non-perturbative parameters (form factor $F^{B M}(q^2=0)$, decay constants $f_M$, light-cone distribution amplitudes $\phi_M$) and perturbative hard-scattering kernels $T_i^{I,II}$, that contain the short-distance dynamics of the flavour-changing quark transition. The latter are currently being worked out to NNLO, i.e.~at $\mathcal{O}(\alpha_s^2)$. While the full set of hard-scattering kernels from spectator scattering ($T_i^{II}$) for tree~\cite{NNLO:T2:tree} and penguin amplitudes~\cite{NNLO:T2:peng} is now available at NNLO, the ones related to the vertex corrections ($T_i^{I}$) are known to date for the tree amplitudes only~\cite{NNLO:T1:tree:GB,NNLO:T1:tree:AC}.
The NNLO calculation is particularly important for direct CP asymmetries that are first generated at $\mathcal{O}(\alpha_s)$. As in any perturbative calculation, it may thus help to reduce scale ambiguities of the leading contribution. The NNLO corrections may even change the pattern of CP asymmetries significantly, since they can potentially be enhanced by large Wilson coefficients (which is not possible at even higher orders since the NNLO terms already have the full complexity).
The current status of the NNLO calculation does not yet allow to discuss CP asymmetries. We may, however, already consider (CP-averaged) branching ratios of tree-dominated decay modes that do not depend significantly on the penguin amplitudes. As these observables are likely to be dominated by their Standard Model contributions, they may serve as important probes for our understanding of the strong-interaction dynamics in non-leptonic decays.
The NNLO analysis of the eleven tree-dominated $B\to\pi\pi/\pi\rho/\rho\rho$ decay modes has been presented in~\cite{Bell:2009fm} (for a similar analysis cf.~\cite{NNLO:T1:tree:AC}). In general colour-allowed decay modes turn out to be under much better theoretical control than colour-suppressed modes. Let us illustrate this point at the amplitude level: for the colour-allowed amplitude in the $\pi\pi$ channels one finds
$\alpha_1(\pi\pi) = 1.013^{+0.023}_{-0.036} + (+0.027^{+0.025}_{-0.022})i$,
which is to be compared with
$\alpha_2(\pi\pi) = 0.195^{+0.134}_{-0.089} + (-0.101^{+0.061}_{-0.063})i$
for the colour-suppressed one. It is striking that the latter suffers from substantial theoretical uncertainties, which can be traced back to a strong (and unfortunate) cancellation between different terms in the perturbative expansion. This makes the real part of $\alpha_2$ particularly sensitive to the spectator scattering mechanism, which is normalized by the hadronic ratio $f_{\pi}\hat{f}_B/\lambda_BF_+^{B \pi}(0)$. The poor knowledge of the $B$ meson parameter $1/\lambda_B=\int_0^\infty d\omega/\omega\; \phi_B(\omega)$, in particular, makes the theoretical prediction of the colour-suppressed amplitude rather uncertain.
In order to test the QCD dynamics in hadronic decays it is useful to consider \emph{ratios} of decay rates rather than absolute branching fractions. Particularly suited are ratios that involve the differential semileptonic decay rate at maximum recoil,
\begin{align}
{\cal{R}}_{M_3}(M_1 M_2) =
\frac{\Gamma(\bar{B} \to M_1 M_2)}
{d\Gamma(\bar{B}^0\to M_3^+\ell^-\bar{\nu}_l)/dq^2|_{q^2=0}}.
\end{align}
Experimentally this requires to measure the semileptonic decay spectrum over a sufficiently large number of $q^2$-bins. Assuming specific parameterizations for the form factor shapes, the spectrum may then be extrapolated to $q^2=0$. At present this information is available for $B\to\pi\ell\nu$ decays~\cite{piellnu}, whereas the data on the $B\to\rho\ell\nu$ spectrum does not yet allow for an accurate extrapolation.
\begin{table*}[b!]
\begin{center}
\begin{tabular}{|r|cccc|} \hline
\hspace*{0cm}&
\hspace*{2cm}&\hspace*{2cm}&\hspace*{2cm}&\hspace*{2cm}
\\[-1.1em]
& ${\cal{R}}_\pi(\pi^-\pi^0)$
& ${\cal{R}}_\pi(\pi^+\pi^-)$
& ${\cal{R}}_\pi(\pi^0\rho^-)$
& ${\cal{R}}_\pi(\pi^+\rho^-)$
\\ [0.1em]
\hline
&&&&
\\[-1.2em]
Theory
& $0.70^{+0.12}_{-0.08}$
& $1.09^{+0.22}_{-0.20}$
& $1.71^{+0.27}_{-0.24}$
& $2.77^{+0.32}_{-0.31}$
\\[0.2em]
Experiment
& $0.81^{+0.14}_{-0.14}$
& $0.80^{+0.13}_{-0.13}$
& $1.57^{+0.32}_{-0.32}$
& $2.43^{+0.47}_{-0.47}$
\\[0.1em]
\hline
\end{tabular}
\end{center}
\vspace{-4.5mm}
\parbox{15.1cm}{\caption{\label{tab:slratios}
Ratios of hadronic and differential semileptonic decay rates in units of $\text{GeV}^2$.}}
\end{table*}
In Table~\ref{tab:slratios} we confront the NNLO prediction of the ${\cal{R}}_{\pi}$-ratios with experimental data. We stress that the theoretical predictions from Table~\ref{tab:slratios} are based on a default set of hadronic input parameters (specified in Table~I of~\cite{Bell:2009fm}), that is motivated by recent lattice and sum rule calculations. We see that the theoretical predictions are in good agreement with the data, which strongly supports the factorization assumption\footnote{The agreement is less pronounced for the ratio ${\cal{R}}_\pi(\pi^+\pi^-)$, which shows a much stronger dependence on the QCD penguin amplitude and hence on the specific input value for the weak phase $\gamma$. This ratio is thus not particularly suited to test the dynamics of the tree amplitudes.}. This is in particular true for the ratio ${\cal{R}}_\pi(\pi^-\pi^0)$, which does not depend on the QCD penguin amplitude and on weak annihilation contributions at all; it thus gives clean access to $|\alpha_1(\pi\pi)+\alpha_2(\pi\pi)|^2$~\cite{facttest}. Taking current data at face value, we may conclude that the colour-suppressed amplitude is somewhat enhanced, which may hint at a lower value of the $B$ meson parameter $\lambda_B\simeq~250$MeV (the default choice adopted in~\cite{Bell:2009fm} is $\lambda_B=(400\pm150)$MeV). It would be interesting to verify if this conclusion is supported by the according ratio in the $\rho$-sector. As long as the experimental information on the semileptonic $B\to\rho\ell\nu$ spectrum is absent, we may instead consider ratios of two hadronic decay rates,
\begin{align}
R(M_1 M_2/M_3 M_4) =
\frac{\Gamma(\bar{B} \to M_1 M_2)}
{\Gamma(\bar{B}' \to M_3 M_4)}.
\end{align}
The ratio $R(\rho_L^- \rho_L^0/\rho_L^+ \rho_L^-) \simeq |\alpha_1(\rho_L\rho_L)+\alpha_2(\rho_L\rho_L)|^2/2|\alpha_1(\rho_L\rho_L)|^2$ yields complementary information on the tree amplitudes from the $\rho$-sector\footnote{The subscript $L$ refers to the longitudinal polarization.}. One should keep in mind, however, that this ratio receives corrections from the QCD penguin amplitude and from weak annihilation in contrast to the semileptonic ratio ${\cal{R}}_{\rho}(\rho_L^- \rho_L^0)$. From the numbers in Table~\ref{tab:hadratios} we infer that the NNLO prediction is again found to be smaller than the experimental value, which supports the hypothesis of enhanced colour-suppressed amplitudes and hence a lower value of $\lambda_B$.
\begin{table*}[b!]
\begin{center}
\begin{tabular}{|r|c|cc|} \hline
\hspace*{0cm}&
\hspace*{2.5cm}&\hspace*{2.5cm}&\hspace*{2.5cm
\\[-1.1em]
& $R(\rho_L^- \rho_L^0/\rho_L^+ \rho_L^-)$
& $R(\pi^0 \rho^0/\pi^0 \pi^0)$
& $R(\pi^0 \rho^0/\rho_L^0 \rho_L^0)$
\\ [0.1em]
\hline
&&
\\[-1.2em]
Theory
& $0.65^{+0.16}_{-0.11}$
& $1.50^{+1.70}_{-1.32}$
& $1.17^{+0.45}_{-0.43}$
\\[0.2em]
Experiment
& $0.89^{+0.14}_{-0.14}$
& $1.29^{+0.36}_{-0.36}$
& $2.90^{+1.45}_{-1.45}$
\\[0.1em]
\hline
\end{tabular}
\end{center}
\vspace{-4mm}
\parbox{15.1cm}{\caption{\label{tab:hadratios}
Ratios of two hadronic decay rates.}}
\end{table*}
Let us finally comment on the colour-suppressed modes, which are more complicated due to their strong dependence on hadronic input parameters. In contrast to the colour-allowed modes, it is in particular \emph{not} possible to reduce these uncertainties by considering semi\-leptonic ${\cal{R}}_{M}$-ratios since their dependence on $|V_{ub}|^2 |F_+^{BM}(0)|^2$ is weak. One may instead try to resolve the correlation among the theoretical uncertainties by considering hadronic ratios of two colour-suppressed modes. As can be seen in~Table~\ref{tab:hadratios}, this does unfortunately not lead to an improvement for the ratio that involves the $\pi^0\pi^0$ decay mode\footnote{Let us emphasize that the agreement between the central values is accidental for this ratio, since the theoretical prediction for the individual branching ratios are quite below the experimental data (the numbers can be found in~\cite{Bell:2009fm}).}. This is different for the ratio $R(\pi^0 \rho^0/\rho_L^0 \rho_L^0)$, which is less contaminated by the QCD penguin amplitudes. Consequently, the dependence on $\lambda_B$ and, somewhat accidentally, the one from the modelled power corrections drop out to a large extent. A more precise experimental value for this ratio may therefore give further insight into the role of power corrections in non-leptonic decays. The dynamics of the colour-suppressed amplitudes, however, should be probed with the cleaner ratios ${\cal{R}}_{\pi}(\pi^- \pi^0)$, ${\cal{R}}_{\rho}(\rho_L^- \rho_L^0)$, ${\cal{R}}_{\rho}(\pi^- \rho^0)$ and ${\cal{R}}_{\pi}(\pi^0\rho^- )$.
\section{Semileptonic B decays}
Semileptonic $b\to u$ decays provide a measure of the CKM matrix element $|V_{ub}|$. The current discrepancy between inclusive and exclusive determinations calls for further progress on both sides. Here we report on the NNLO calculation for inclusive $B\to X_u\ell\nu$ decays in the so-called BLNP approach~\cite{BLNP}.
The theoretical description of inclusive $B\to X_u\ell\nu$ decays is complicated by the fact that experimental measurements have to introduce kinematical cuts to suppress the $B\to X_c\ell\nu$ background. This restricts the experimental information to the shape-function region in which the hadronic final state has large energy $E_X \sim m_b$ but moderate invariant mass $p_X^2 \sim m_b \Lambda_{QCD}$. In this region of phase space a factorization formula for the structure functions has been put forward \cite{Korchemsky:1994jb}
\begin{eqnarray}
W_i &\;\simeq\;&
H_{i} \,
\int d\omega \; J(p_\omega^2) \; S(\omega),
\end{eqnarray}
which contains an universal non-perturbative quantity, the shape function $S$, and two perturbatively calculable objects, hard coefficient functions $H_{i}$ and a jet function $J$, that encode the short-distance effects. The NNLO calculation of the latter is now complete. While the two-loop corrections to the jet function have been worked out in~\cite{Becher:2006qw}, the hard coefficient functions have recently been computed to NNLO by various groups~\cite{BtoXuellnu}. The latter calculation required to match the flavour-changing $V-A$ current from QCD onto soft-collinear effective theory~\cite{SCET}. It has further been generalized to the tensor current, which finds applications in electroweak penguin decays~\cite{BBHL}.
The numerical impact of the NNLO corrections on the inclusive determination of $|V_{ub}|$ has recently been analyzed in~\cite{Greub:2009sv}. Starting from the two-loop expressions for the hard and jet functions, the authors implemented the renormalization group improvement and a specific model for the shape function. From their analysis of partial decay rates, they conclude that the NNLO corrections can be significant. This statement, however, depends on the choice of the (arbitrary) matching scale $\mu_i\sim(m_b \Lambda_{QCD})^{1/2}$. For $\mu_i= 1.5$~GeV, which was the default choice in the earlier BLNP analysis~\cite{BLNP}, the NNLO corrections are found to be important; they typically lower partial decay rates by about $15-20\%$ while at the same time reducing the perturbative uncertainties.
\begin{table*}[b!]
\begin{center}
\begin{tabular}{|lccc|} \hline
\hspace*{0cm}&
\hspace*{2.5cm}&\hspace*{2.5cm}&\hspace*{2.5cm
\\[-1.1em]
Exp. & Method
& $|V_{ub}|~[10^{-3}]$
& $|V_{ub}|~[10^{-3}]$
\\[-0.1em]
&
& NLO
& NNLO
\\ [0.1em]
\hline
&&
\\[-1.2em]
BABAR
& $E_l>2.0$~GeV
& $3.97\pm0.22^{+0.37+0.26}_{-0.23-0.25}$
& $4.30\pm0.24^{+0.26+0.28}_{-0.20-0.27}$
\\[0.2em]
BELLE
& $M_X<1.7$~GeV
& $3.55\pm0.24^{+0.22+0.21}_{-0.13-0.19}$
& $3.87\pm0.26^{+0.21+0.21}_{-0.13-0.19}$
\\[0.2em]
BABAR
& $P_+<0.66$~GeV
& $3.30\pm0.23^{+0.27+0.25}_{-0.16-0.22}$
& $3.55\pm0.24^{+0.19+0.24}_{-0.13-0.21}$
\\[0.1em]
\hline
\end{tabular}
\parbox{12.1cm}{\caption{\label{tab:Vub}
Values of $|V_{ub}|$ deduced from different experimental measurements of partial $B\to X_u\ell\nu$ decay rates (the numbers are taken from~\cite{Greub:2009sv}).}}
\end{center}
\end{table*}
In their determination of $|V_{ub}|$ the authors combine the NNLO prediction of the leading term in the heavy quark expansion with known power corrections up to $\mathcal{O}(1/m_b^2)$. From the experimental information on partial decay rates, which are based on different types of experimental cuts (lepton energy $E_l$, hadronic invariant mass $M_X$, hadronic variable $P_+ = E_X - |\vec{P}_X|$), they deduce sample values for $|V_{ub}|$. Some of their results are collected in Table~\ref{tab:Vub}, which illustrate that the central values are shifted significantly at NNLO. Concerning the error estimate, the first one reflects the experimental uncertainty, while the improvement on the perturbative uncertainty can be seen in the second one. One further infers from the last error that the numerical value of the $b$-quark mass, which enters certain moment constraints of the shape-function model, has a large impact on the determination of $|V_{ub}|$. Given that the NNLO calculation has increased the discrepancy with the exclusive determination, further theoretical progress on the treatment of shape function effects is highly desirable (cf., for instance,~\cite{Ligeti:2008ac} for an alternative implementation that is not based on a specific model).
|
\section{Introduction}
Yurii Fedorovich Smirnov passed away in 2008 and Marcos Moshinsky in 2009. They were
two famous physicists with common interests in nuclear physics, atomic and molecular
physics, and mathematical physics. More generally, both of them were at the origin of
significant achievements in symmetry methods in physics. They actively participated in
several Symmetries in Science Symposia in Bregenz. These two giants had parallel
centers of interest in the sense that they developed separately some complementary
works in nuclear physics (that led in particular to the concept of Moshinsky-Smirnov coefficients),
dealt with some related problems in atomic, molecular and mathematical physics,
and, finally, combined efforts to produce a beautiful and very useful book \cite{MosSmi}
on the applications of the harmonic oscillator system to various areas of physics and chemistry.
It is not the purpose of these notes to extensively list and analyse the numerous
papers by Marcos and Yurii. I shall focus on some particular facets of their works. I
had the opportunity to meet Marcos and Yurii several times in Bregenz and on several
other occasions, and to discuss with them about Wigner-Racah algebras for finite groups, Lie groups
and quantum groups. I had also a chance to collaborate with Yurii Smirnov. Therefore, I shall devote
the main part of these notes to some specific domains of importance to Yurii and Marcos
and to some more personal reminiscences on Yurii.
\section{Marcos Moshinsky}
Marcos Moshinsky was a Mexican physicist. He was born in Kiev (Ukraine) in 1921. He
arrived as a refugee in Mexico when he was three years old and obtained
Mexican citizenship in 1942. He received a Bachelor's degree in physics from
the {\it Universidad Nacional Aut\'onoma de M\'exico} (U.N.A.M.) and a Ph.D. degree in
theoretical physics, under the guidance of Eugene P. Wigner, from Princeton
University. Marcos was also the recipient a post-doctoral fellowship at the {\it Institut Henri Poincar\'e}
in Paris. Afterwards, he returned to Mexico and pursued a brillant career at the U.N.A.M. in Mexico City.
Professor Marcos Moshinsky had important responsabilities as the President of
the {\it Sociedad Mexicana de F\'\i sica}, as a member of {\it El Colegio Nacional},
and as a member of the editorial board of several international scientific reviews. He
produced and/or co-produced more than 200 scientific papers and four books among which the most
well-known are the one written in collaboration with Thomas A. Brody on transformation brackets
for nuclear shell-model calculations \cite{BroMos} and the one with Yurii F. Smirnov on the
applications of the harmonic oscillator in various fields of physics and quantum
chemistry \cite{MosSmi}. He received several prizes, namely the
{\it Premio Nacional de Ciencias y Artes} in 1968,
{\it Premio Luis Elizondo} in 1971,
{\it Premio U.N.A.M. de Ciencias Exactas} in 1985,
{\it Premio Pr\'\i ncipe de Asturias de Investigaci\'on Cient\'\i fica y T\'ecnica} in 1988, and
the prestigious UNESCO Science Prize in 1997 for his work in nuclear physics. He also received
the Wigner medal in 1998.
\section{Marcos and Yurii}
A first seminal paper by Marcos concerned the transient dynamics of particle wavefunctions, a phenomenon that
gives rise to diffraction in time \cite{Mos1952}. However, most of his scientific work dealt with collective
models of the nucleus, canonical transformations in quantum mechanics, and group theoretical methods in physics,
with a special emphasis on symplectic symmetry in nuclear, atomic and molecular physics. These themes were
of interest to Yurii too. Following the pioneering work of Talmi (who prepared his Ms.S. thesis with Guilio Racah,
his Doctorate thesis with Wolfgang Pauli and who was a post-doctoral fellow with Eugene P. Wigner) \cite{Talmi1952},
both Marcos and Yurii were
interested in the description of pairs of nucleons in a harmonic-oscillator potential. In 1959, Moshinsky developed a
formalism to connect wavefunctions in two different coordinate systems for two particles (with identical masses)
in a harmonic-oscillator potential \cite{Mos1959}. In this formalism, any two-particle wavefunction
$|n_1 \ell_1, n_2 \ell_2, \lambda \mu \rangle$, expressed in coordinates with respect to the
origin of the harmonic-oscillator potential, is a linear combination of wavefunctions
$|n \ell, N L, \lambda \mu \rangle$, expressed in relative and centre-of-mass coordinates of the two particles. The so-called
transformation brackets $\langle n \ell, N L, \lambda | n_1 \ell_1, n_2 \ell_2, \lambda \rangle$ make it possible to pass
from one coordinate system to the other. Moshinsky gave an explicit expression of these coefficients in the case
$n_1 = n_2 = 0$ and derived recurrence relations that can be used to obtain the coefficients for
$n_1 \not= 0$ and $n_2 \not= 0$ from those for $n_1 = n_2 = 0$ \cite{Mos1959}. Along this vein, Brody and Moshinsky
published extensive tables of transformation brackets \cite{BroMos}. At the end of the fifties, Smirnov worked out a
similar problem, viz. the calculation of the Talmi coefficients for unequal mass nucleons, and gave solution for
the case $n_1 \not= 0$ and $n_2 \not= 0$ \cite{Smi1961, Smi1962}. (Indeed, the transformation brackets and the Talmi
coefficients are connected via a double Clebsch-Gordan transformation.) The coefficients
$\langle n \ell, N L, \lambda | n_1 \ell_1, n_2 \ell_2, \lambda \rangle$, called {\it transformation brackets} by Moshinsky
and {\it total Talmi coefficients} by Smirnov, are now referred to as Moshinsky-Smirnov coefficients. Both the
Moshinsky-Smirnov coefficients and the Talmi coefficients were revisited at the end of the seventies in terms of
generating functions in the framework of the approaches of Julian S. Schwinger and Valentine Bargmann to the
harmonic-oscillator bases \cite{Mehdi1980}. (The work by Mehdi Hage Hassan \cite{Mehdi1980}, who prepared his
State Doctorate thesis at the {\it Institut de Physique Nucl\'eaire de Lyon} and conducted his career in Beyrouth, constitutes a very deep and original
approach to the Talmi coefficients and Moshinsky-Smirnov coefficients.) It should be noted that the transformation brackets
or Moshinsky-Smirnov coefficients are also of importance for atoms and molecules as shown by Marcos and Yurii in their book \cite{MosSmi}
written during the time Yurii was a visiting professor at the {\it Instituto de F\'\i sica} of the
{\it Universidad Nacional Aut\'onoma de M\'exico}.
A second area of common interest to both Marcos and Yurii concerns the many-body
problem considered from the point of view of unitary and symplectic groups and the use of nonlinear and
nonbijective canonical transformations. In this vein, Moshinsky and some of his collaborators
introduced the concept of an ambiguity group, a group required for taking into account the nonbijectivity
of certain canonical transformations \cite{nonbijective}. Indeed, this concept is closely related to the
one of Lie algebra under constraint \cite{KibWin}, which in turn is connected to nonbijective transformations
like Hopf fibrations on spheres and Hopf fibrations on hyperbolids \cite{Lambert}. Among the nonbijective
transformations, one may mention the ${\bf R}^4 \to {\bf R}^3$ Kustaanheimo-Stiefel transformation
and the ${\bf R}^2 \to {\bf R}^2$ Levi-Civita transformation as well as their various extensions
\cite{Lambert, MedhiKib}. In particular, the Kustaanheimo-Stiefel transformation allows one to pass from
a four-dimensional harmonic oscillator subjected to a constraint to the three-dimensional hydrogen atom
(see for instance \cite{KibNeg}). This subject was of interest to Yurii and he revisited the hydrogen-oscillator
connection with Corrado Campigotto \cite{CamSmi}.
The harmonic oscillator is a central ingredient in numerous studies
by Smirnov and Moshinsky. Many applications of the nonrelativistic and relativistic harmonic oscillators
to modern physics (from molecules, atoms and nuclei to quarks) were pedagogically exposed in the book by
Marcos and Yurii \cite{MosSmi} with a special attention paid to the $n$-body problem (in the Hartree-Fock
approximation), the nuclear collective motion and the interacting boson model.
But their common interests were not limited to transformation brackets and harmonic oscillators. Let us briefly mention that
both of them were also interested in group theoretical methods and symmetry methods in physics and also contributed to several
fields of mathematical physics including, for instance, the state labelling problem, special functions, and generating functions
(see, for example, \cite{MPSW, RRS}).
\section{Yurii Fedorovich Smirnov}
Yurii Fedorovich Smirnov was a Russian physicist. He was born in the city of Il'inskoe
(Yaroslavl' region, Russia) in 1935. He graduated from Moscow State University. Subsequently, he
completed his Doctorate thesis at the same university under the guidance of Yurii M. Shirokov and
benefited from fruitful contacts with other distinguished physicists including Yakov A. Smorodinsky. He
pursued his career in the (Skobeltsyn) Institute of Nuclear Physics and in the Physics Department of
(Lomonosov) Moscow State University with many stays abroad. The
last fifteen years of his life were shared between Moscow and Mexico City where he was a visiting professor and later
a professor at the {\it Instituto de F\'\i sica} and at the {\it Instituto de
Ciencias Nucleares} of the U.N.A.M. (he spent almost 11 years in Mexico). He received prestigious awards:
the K.D. Sinel'nikov Prize of the Ukrainian Academy of Sciences in 1982 and the M.V. Lomonosov Prize in 2002. He was
also a member of the Academy of Sciences of Mexico.
Yurii Smirnov authored and/or co-authored eleven books and more than 250 scientific articles. He
also translated several scientific books into Russian. He translated, for example, a book on the
harmonic oscillator written by Marcos Moshinsky in 1969, precisely the book that was a starting point for
their common book on the same subject, published in 1996 \cite{MosSmi}. He was a member of the editorial
board of several journals and a councillor of the scientific councils of the Skobeltsyn Institute of Nuclear
Physics and of the Chemistry Department of Moscow State University, as well as of the Institute for
Theoretical and Experimental Physics (ITEP) in Moscow.
\section{Some personal reminiscences on Yurii}
My first contact with the work of Yurii Smirnov dates back to 1978 when my colleague
J. Patera showed me, on the occasion of a NATO Advanced Study Institute organised in Canada by J.C. Donini,
a beautiful book written by D.T. Sviridov and Yu.F. Smirnov \cite{Smirnov1}. This book
dealt with the spectroscopy of $d^N$ ions in inhomogeneous electric fields (part of a
disciplinary domain known as crystal- and ligand-field theory in condensed matter physics and
explored via the theory of level splitting from a theoretical point of view). In 1979, B.I. Zhilinski\u\i ,
while visiting Dijon and Lyon in France in the context of an exchange programme between
the USSR and France, gave me another interesting book, on $f^N$ ions in
crystalline fields, written by D.T. Sviridov, Yu.F. Smirnov and V.N. Tolstoy \cite{Smirnov2}. At
that time, the references for mathematical aspects of crystal- and ligand-field theory were based
on works by Y. Tanabe, S. Sugano and H. Kamimura from Japan \cite{Tanabe}, J.S. Griffith from
England \cite{Griffith}, and Tang Au-chin and his collaborators from China \cite{Tang} (see also
some contributions by the present author \cite{KibJMSIJQC}). The two above-mentioned books
by Smirnov and his colleagues shed some new light on the mathematical analysis of spectroscopic
and magnetic properties of partly filled shell ions in molecular and crystal surroundings. In
particular, special emphasis was put on the derivation of the Wigner-Racah algebra of a finite
group of molecular and crystallographic interest from that of the group $SO(3) \sim SU(2)/Z_2$.
My second (indirect) contact with Yurii is related to an invitation to participate in the fifth workshop on
{\it Symmetry Methods in Physics} in Obninsk in July 1991. Unfortunately, I did not get my visa on time
reducing my participation to a paper in the proceedings of the workshop edited by Yu.F. Smirnov and
R.M. Asherova \cite{KiblerObninsk}.
In the beginning of the 1990's, I had a chance to discover another facet of Yurii's work. In 1989,
a Russian speaking student from Switzerland, C. Campigotto, spent one year in the group of Prof.~Smirnov. He
started working on the Kustaanheimo-Stiefel transformation, an ${\bf R}^4 \to {\bf R}^3$
transformation associated with the Hopf fibration ${S}^3 \to {S}^2$ with compact fiber $S^1$. (Such a transformation
makes it possible to connect the Kepler-Coulomb system in ${\bf R}^3$ to the isotropic harmonic oscillator
in ${\bf R}^4$.) Then, Campigotto (well-prepared by Smirnov and his team, especially Andrey M. Shirokov and Valeriy N. Tolstoy)
came to Lyon to prepare a Doctorate thesis (partly published in \cite{Campigotto}). He defended his thesis in 1993
with George S. Pogosyan (representing Yu.F. Smirnov) as a member of the jury.
A fourth opportunity to work with Yurii stemmed from our mutual interest in quantum groups
and in nuclear and atomic spectroscopy. I meet him for the first time in Dubna in 1992. We then
started a collaboration (partly with R.M. Asherova) on $q$- and $qp$-boson calculus in the
framework of Hopf algebras associated with the Lie algebras $su(2)$ and $su(1,1)$ \cite{KibSmiCamAsh}. In addition,
we pursued a group-theoretical study of the Coulomb energy averaged over the $n \ell^N$--atomic
states with a definite spin \cite{IJQC}. We also had fruitful exchanges in nuclear physics. Indeed, Prof.~Smirnov
and his colleagues D. Bonatsos (from Greece), S.B. Drenska, P.P. Raychev and R.P. Roussev (all from Bulgaria)
developed a model based on a one-parameter deformation of $SU(2)$ for dealing with rotational bands
of deformed nuclei and rotational spectra of molecules \cite{Bonatsos} (see also \cite{Georgieva}). Along
the same line, a student of mine, R. Barbier, developed in his thesis a
two-parameter deformation of $SU(2)$ with application to superdeformed nuclei in mass region $A \sim 130-150$
and $A \sim 190$ (partly published in \cite{Barbier}). It was a real pleasure to receive Yurii in Lyon on the
occasion of the defence of the Barbier thesis in 1995. Indeed, from 1992 to 1995, Yurii made four stays in
Lyon (one with his wife Rita and one with his daughter Tatyana) and we jointly participated in several meetings,
one in Clausthal in Germany (organised by H.-D. Doebner, V.K. Dobrev and A.G. Ushveridze) and two in Bregenz in
Austria (organised by B. Gruber and M. Ramek).
I cannot do justice to all of the fields in which
Yurii was recognized as a superb researcher. It is enough to say that he
contributed to many domains of mathematical physics (e.g., finite groups
embedded in compact or locally compact groups, Lie groups and Lie algebras,
quantum groups, special functions) and theoretical physics (e.g., nuclear,
atomic and molecular physics, crystal- and ligand-field theory). Let me
mention, among other fields, that he achieved alone and with collaborators
significant advances in the theory of clustering of shell-model (nuclear) systems
\cite{clustering}, in projection operator techniques for simple Lie groups \cite{Tolstoy}, in the
theory of heavy ion collisions \cite{SmiTch}, and in the so-called $J$-matrix formalism for quantum
scattering theory (see \cite{Cocoyoc} and references therein). The $J$-matrix formalism requires the
solution of three-term recurrence relations (or second-order difference equations);
along this line, Yurii and some of his collaborators
published several works (see for instance \cite{Shirokov}). As another major contribution,
at the end of the sixties he proposed in collaboration with Vladimir G. Neudatchin a method, the
so-called (e,2e) method (an analog of the (p,2p) method used in nuclear
physics), for the experimental investigation of the electronic structure
of atoms, molecules and solids; this method was successfully tested in
many laboratories around the world (see \cite{Neudatchin} and references therein).
Yurii was also an exceptional teacher. It was very pleasant, profitable and inspiring
to be taught by him. I personally greatly benefited from discussions with
Yurii Smirnov.
\section*{Closing}
Yurii and Marcos had many students who are now famous physicists. They interacted with many
collaborators in their countries and abroad, and had an influence on many scientists. Marcos
Moshinsky and Yurii Fedorovich Smirnov will remain examples for many of us. We shall not
forget them.
\section*{References}
|
\section{Introduction}\label{s1}
Since Borg--Marchenko-type uniqueness theorems were first formulated in the context of scalar Schr\"odinger operators on half-lines, we start with a brief review of these results: Let $H_j = -\frac{d^2}{dx^2} + V_j$, $V_j\in L^1 ([0,R]; dx)$ for all $R>0$, $V_j$
real-valued, $j=1,2$, be two self-adjoint operators in $L^2 ([0,\infty); dx)$ which, just for simplicity, have a Dirichlet boundary condition at $x=0$ (and possibly a self-adjoint boundary condition at infinity). Let $m_j(z)$, $z\in{\mathbb{C}}\backslash{\mathbb{R}}$, be the
Weyl--Titchmarsh $m$-functions associated with $H_j$, $j=1,2$. Then the celebrated Borg--Marchenko uniqueness theorem, in this particular context, reads as follows:
\begin{theorem} \label{t1.1}
Suppose
\begin{equation}
m_1(z) = m_2(z), \;\; z\in{\mathbb{C}}\backslash{\mathbb{R}}, \, \text{ then } \,
V_1(x) = V_2(x) \, \text{ for a.e.\ $x\in [0,\infty)$.} \label{1.1}
\end{equation}
\end{theorem}
This result was published by Marchenko \cite{Ma50} in 1950. Marchenko's extensive treatise on spectral theory of one-dimensional Schr\"odinger operators \cite{Ma52}, repeating the proof of his uniqueness theorem, then appeared in 1952, which also marked the appearance of Borg's proof of the uniqueness theorem \cite{Bo52} (apparently, based on his lecture at the 11th Scandinavian Congress of Mathematicians held at Trondheim, Norway in 1949).
We emphasize that Borg and Marchenko also treat the general case of
non-Dirichlet boundary conditions at $x=0$ (in which equality of the two $m$-functions also identifies the two boundary conditions), moreover, Marchenko also simultaneously discussed the half-line and the finite interval case. For brevity we chose to illustrate the simplest possible case only.
To the best of our knowledge, the only alternative approaches to Theorem \ref{t1.1}
are based on the Gelfand--Levitan solution \cite{GL51} of the inverse spectral problem published in 1951 (see also Levitan and Gasymov \cite{LG64}) and alternative variants due to M.~Krein \cite{Kr51}, \cite{Kr53}. For over 45 years, Theorem \ref{t1.1} stood the test of time and resisted any improvements. Finally, in 1998, Simon \cite{Si98} proved the following spectacular result, a local Borg--Marchenko theorem (see part $(i)$ below) and a significant improvement of the original Borg--Marchenko theorem (see part $(ii)$ below):
\begin{theorem}\label{t1.2} ${}$ \\
$(i)$ Let $a>0$, $0<\varepsilon<\pi/2$ and
suppose that
\begin{equation} \label{1.4}
|m_1 (z) - m_2(z)| \underset{|z|\to\infty}{=} O(e^{-2\Im (z^{1/2})a})
\end{equation}
along the ray $\arg(z) = \pi-\varepsilon$. Then
\begin{equation} \label{1.5}
V_1(x) = V_2 (x) \, \text{ for a.e.\ $x\in [0,a]$.}
\end{equation}
$(ii)$ Let $0<\varepsilon <\pi/2$ and suppose
that for all $a>0$,
\begin{equation}
|m_1(z) - m_2(z)| \underset{|z|\to\infty}{=} O(e^{-2\Im (z^{1/2})a}) \label{1.6}
\end{equation}
along the ray $\arg(z) = \pi -\varepsilon$. Then
\begin{equation}
V_1(x) = V_2(x) \, \text{ for a.e.\ $x\in [0,\infty)$.} \label{1.7}
\end{equation}
\end{theorem}
The ray $\arg(z) = \pi -\varepsilon$, $0<\varepsilon < \pi/2$ chosen in
Theorem \ref{t1.2} is of no particular importance. A limit taken along any non-self-intersecting curve ${\mathcal C}$ going to infinity in the sector
$\arg(z)\in ((\pi/2)+\varepsilon, \pi -\varepsilon)$ is permissible. For simplicity we only discussed the Dirichlet boundary condition $u(0)=0$ thus far. However, everything extends to the case of general boundary conditions $u'(0) + h u(0) = 0$, $h\in{\mathbb{R}}$. Moreover, the case of a finite interval problem on $[0,b]$, $b\in (0,\infty)$, instead of the half-line $[0,\infty)$ in Theorem \ref{t1.2}\,$(i)$, with $0<a<b$, and a self-adjoint boundary condition at $x=b$ of the type $u'(b) + h_b u(b) = 0$, $h_b \in {\mathbb{R}}$, can be handled as well. All of this is treated in detail in \cite{GS00a}.
Remarkably enough, the local Borg--Marchenko theorem proven by Simon \cite{Si98} was just a by-product of his new approach to inverse spectral theory for half-line
Schr\"odinger operators. Actually, Simon's original result in \cite{Si98} was obtained under a bit weaker conditions on $V$; the result as stated in Theorem \ref{t1.2} is taken from \cite{GS00a} (see also \cite{GS00}). While the original proof of the local
Borg--Marchenko theorem in \cite{Si98} relied on the full power of a new formalism in inverse spectral theory, a short and fairly elementary proof of Theorem \ref{t1.2} was presented in \cite{GS00a}. Without going into further details at this point, we also mention that \cite{GS00a} contains the analog of the local Borg--Marchenko uniqueness result, Theorem \ref{t1.2} for Schr\"odinger operators on the real line. In addition, the case of half-line Jacobi operators and half-line matrix-valued Schr\"odinger operators was dealt with in \cite{GS00a}.
We should also mention some work of Ramm \cite{Ra99}, \cite{Ra00}, who provided a proof of Theorem~\ref{t1.2}\,$(i)$ under the additional assumption that $V_j$ are
short-range potentials satisfying $V_j\in L^1([0,\infty); (1+|x|)dx)$, $j=1,2$.
A very short proof of Theorem \ref{t1.2}, close in spirit to Borg's original paper
\cite{Bo52}, was subsequently found by Bennewitz \cite{Be01}. Still other proofs were presented in \cite{Ho01} and \cite{Kn01}. Various local and global uniqueness results for matrix-valued Schr\"odinger, Dirac-type, and Jacobi operators were considered in
\cite{CG02}, \cite{FKRS07}, \cite{GKM02}, \cite{Sa90}, \cite{Sa02}, and \cite{Sa06}. A local
Borg--Marchenko theorem for complex-valued potentials has been proved in
\cite{BPW02}; the case of semi-infinite Jacobi operators with complex-valued coefficients was studied in \cite{We04}. This circle of ideas has been reviewed in \cite{Ge07}.
After this review of Borg--Marchenko-type uniqueness results for Schr\"odinger operators, we now turn to the principal object of our interest in this paper, the so-called CMV operators. CMV operators are a special class of unitary semi-infinite five-diagonal matrices. But for simplicity, we confine ourselves in this introduction to a discussion of CMV operators on ${\mathbb{Z}}$, that is, doubly infinite CMV operators. Let $\alpha$ be a sequence of $m\times m$ matrices, $m\in{\mathbb{N}}$, with entries in ${\mathbb{C}}$,
$\alpha=\{\alpha_k\}_{k \in {\mathbb{Z}}}$ such that $\|\alpha_k \|_{{\mathbb{C}}^{m\times m}} < 1$,
$k\in{\mathbb{Z}}$. The unitary operator ${\mathbb{U}}$ on $\ltm{{\mathbb{Z}}}$ then can be written as a special five-diagonal doubly infinite matrix in the standard basis of $\ell^2({\mathbb{Z}})^m$ as in
\eqref{3.18}. For the corresponding half-lattice CMV operators ${\mathbb{U}}_{+,k_0}$, in
$\ell^2([k_0,\infty)\cap{\mathbb{Z}})^m$ we refer to \eqref{3.33} and \eqref{3.34}.
The actual history of CMV operators (with scalar coefficients $\alpha_k \in{\mathbb{C}}$, $k\in{\mathbb{Z}}$) is quite interesting: The corresponding unitary semi-infinite five-diagonal matrices were first introduced in 1991 by
Bunse--Gerstner and Elsner \cite{BGE91}, and subsequently discussed in detail by Watkins \cite{Wa93} in 1993 (cf.\ the recent discussion in Simon \cite{Si06}). They were subsequently rediscovered by Cantero, Moral, and Vel\'azquez (CMV) in \cite{CMV03}. In \cite[Sects.\ 4.5, 10.5]{Si04}, Simon introduced the corresponding notion of unitary doubly infinite five-diagonal matrices and coined the term ``extended'' CMV matrices. For simplicity, we will just speak of CMV operators whether or not they are half-lattice or full-lattice operators. We also note that in a context different from orthogonal polynomials on the unit circle, Bourget, Howland, and Joye \cite{BHJ03} introduced a family of doubly infinite matrices with three sets of parameters which, for special choices of the parameters, reduces to two-sided CMV matrices on ${\mathbb{Z}}$. Moreover, it is possible to connect unitary block Jacobi matrices to the trigonometric moment problem (and hence to CMV matrices) as discussed by Berezansky and Dudkin \cite{BD05}, \cite{BD06}.
The relevance of this unitary operator ${\mathbb{U}}$ on $\ell^2({\mathbb{Z}})^m$, more precisely, the relevance of the corresponding half-lattice CMV operator ${\mathbb{U}}_{+,0}$ in
$\ell^2({\mathbb{N}}_0)^m$ is derived from its intimate relationship with the trigonometric moment problem and hence with finite measures on the unit circle ${\partial\hspace*{.2mm}\mathbb{D}}$. (Here
${\mathbb{N}}_0={\mathbb{N}}\cup\{0\}$.)
This will be reviewed in some detail in Section \ref{s3} but we also refer to the
monumental two-volume treatise by Simon \cite{Si04} (see also \cite{Si04b} and
\cite{Si05}) and the exhaustive bibliography therein. For classical results on orthogonal polynomials on the unit circle we refer, for instance, to \cite{Ak65},
\cite{Ge46}--\cite{Ge61}, \cite{Kr45}, \cite{Sz20}--\cite{Sz78},
\cite{Ve35}, \cite{Ve36}. More recent references relevant to the spectral theoretic content of this paper are \cite{De07}, \cite{GJ96}--\cite{GT94},
\cite{GZ06}, \cite{GZ06a}, \cite{GN01}, \cite{PY04}, and \cite{Si04a}. The full-lattice CMV operators ${\mathbb{U}}$ on ${\mathbb{Z}}$ are closely related to an important, and only recently intensively studied, completely integrable nonabelian version of the defocusing nonlinear
Schr\"odinger equation (continuous in time but discrete in space), a special case of the Ablowitz--Ladik system. Relevant references in this context are, for instance,
\cite{AL75}--\cite{APT04}, \cite{GGH05}, \cite{GH05}--\cite{GHMT07a}, \cite{Li05},
\cite{MEKL95}--\cite{Ne06}, \cite{Sc89}, \cite{Ve99}, and the literature cited therein. We emphasize that the case of matrix-valued coefficients $\alpha_k$ is considerably less studied than the case of scalar coefficients.
We note that our discussion of CMV operators will be undertaken in the spirit of
\cite{GKM02}, where (local and global) uniqueness theorems for full-line (resp.,
full-lattice) problems are formulated in terms of diagonal Green's matrices $g(z,x_0)$ and their $x$-derivatives $g^\prime (z,x_0)$ at some fixed $x_0\in{\mathbb{R}}$, for matrix-valued Schr\"odinger and Dirac-type operators on ${\mathbb{R}}$ and similarly for matrix-valued Jacobi operators on ${\mathbb{Z}}$. While we prove half-lattice and full-latice uniqueness results in our principal Section \ref{s5}, we now confine ourselves in this introduction to just two typical results for CMV operators on ${\mathbb{Z}}$ with matrix-valued coefficients:
We use the following notation for the diagonal and for the neighboring off-diagonal entries of the Green's matrix of ${\mathbb{U}}$ (i.e., the discrete integral kernel of $({\mathbb{U}}-zI)^{-1}$),
\begin{align}
g(z,k) = ({\mathbb{U}}-Iz)^{-1}(k,k), \quad
h(z,k) = \begin{cases}
({\mathbb{U}}-Iz)^{-1}(k-1,k), & k \text{ odd}, \\
({\mathbb{U}}-Iz)^{-1}(k,k-1), & k \text{ even},
\end{cases}\quad k\in{\mathbb{Z}},\; z\in\mathbb{D}. \label{1.9}
\end{align}
The next uniqueness result then holds for the full-lattice CMV operator ${\mathbb{U}}$.
\begin{theorem} \label{t1.3}
Let $m\in{\mathbb{N}}$ and assume $\alpha=\{\alpha_k\}_{k\in{\mathbb{Z}}}$ be a sequence of
$m \times m$ matrices with complex entries such that
$\norm{\alpha_k}_{{\mathbb{C}^{m\times m}}} < 1$ and let $k_0\in{\mathbb{Z}}$. Then any of the
following two sets of data
\begin{enumerate}[$(i)$]
\item $g(z,k_0)$ and $h(z,k_0)$ for all $z$ in a sufficiently small neighborhood
of the origin under the assumption that $h(0,k_0)$ is invertible;
\item $g(z,k_0-1)$ and $g(z,k_0)$ for all $z$ in a sufficiently small neighborhood
of the origin and $\alpha_{k_0}$ under the assumption $\alpha_{k_0}$ is invertible;
\end{enumerate}
uniquely determine the matrix-valued Verblunsky coefficients $\{\alpha_k\}_{k\in{\mathbb{Z}}}$, and
hence the full-lattice CMV operator ${\mathbb{U}}$ defined in \eqref{3.18}.
\end{theorem}
In the subsequent local uniqueness result, $g^{(j)}$ and $h^{(j)}$ denote the
corresponding quantities in \eqref{1.9} associated with the
matrix-valued Verblunsky coefficients $\alpha^{(j)}$, $j=1,2$.
\begin{theorem} \label{t1.4}
Let $m\in{\mathbb{N}}$ and assume $\alpha^{(\ell)}=\{\alpha_k^{(\ell)}\}_{k\in{\mathbb{Z}}}$ be sequences of
$m \times m$ matrices with complex entries such that
$\norm{\alpha_k^{(\ell)}}_{{\mathbb{C}^{m\times m}}} < 1$, $k\in{\mathbb{Z}}$, $\ell=1,2$. Moreover, assume
$k_0\in{\mathbb{Z}}$, $N\in{\mathbb{N}}$. Then for the full-lattice
problems associated with $\alpha^{(1)}$ and $\alpha^{(2)}$ the following
local uniqueness results hold:
\begin{enumerate}[$(i)$]
\item
If either $h^{(1)}(0,k_0)$ or $h^{(2)}(0,k_0)$ is invertible and
\begin{align}
\begin{split}
&\big\|g^{(1)}(z,k_0)-g^{(2)}(z,k_0)\big\|_{\mathbb{C}^{m\times m}}
+ \big\|h^{(1)}(z,k_0)-h^{(2)}(z,k_0)\big\|_{\mathbb{C}^{m\times m}} \underset{z\to 0}{=} o(z^N), \label{1.10} \\
& \, \text{then } \, \alpha^{(1)}_k = \alpha^{(2)}_k \,\text{ for }\,
k_0-N \leq k\leq k_0+N+1.
\end{split}
\end{align}
\item
If $\alpha^{(1)}_{k_0}=\alpha^{(2)}_{k_0}$, $\alpha^{(1)}_{k_0}$ is
invertible, and
\begin{align}
\begin{split}
&\big\|g^{(1)}(z,k_0-1)-g^{(2)}(z,k_0-1)\big\|_{\mathbb{C}^{m\times m}} +
\big\|g^{(1)}(z,k_0)-g^{(2)}(z,k_0)\big\|_{\mathbb{C}^{m\times m}} \underset{z\to 0}{=} o(z^N), \label{1.11} \\
& \, \text{then } \, \alpha^{(1)}_k = \alpha^{(2)}_k \,\text{ for }\,
k_0-N-1 \leq k\leq k_0+N+1.
\end{split}
\end{align}
\end{enumerate}
\end{theorem}
The special case of CMV operators with scalar Verblunsky coefficients has recently been discussed in \cite{CGZ07}.
Finally, a brief description of the content of each section in this paper: In Section \ref{s3} we develop the basic Weyl--Titchmarsh theory for half-lattice CMV operators with matrix-valued Verblunsky coefficients. The analogous theory for full-line CMV operators is developed in Section \ref{s4}. Weyl--Titchmarsh theory for CMV operators with matrix-valued Verblunsky coefficients is a subject of independent interest and of fundamental importance in the remainder of this paper. Section \ref{s5} contains our new Borg--Marchenko-type uniqueness results for half-lattice and full-lattice CMV operators with matrix-valued Verblunsky coefficients. Appendix \ref{sA} summarizes basic facts on matrix-valued Caratheodory and Schur functions relevant to this paper.
\section{Weyl--Titchmarsh Theory for Half-Lattice CMV Operators \\ with
Matrix-Valued Verblunsky Coefficients}
\label{s3}
In this section we present the basics of Weyl--Titchmarsh theory for half-lattice CMV operators with matrix-valued Verblunsky coefficients. We closely follow the corresponding treatment of scalar-valued Verblunsky coefficients in \cite{GZ06}.
We should note that while there is an extensive literature on orthogonal matrix-valued polynomials on the real line and on the unit circle, we refer, for instance, to
\cite{AN84}, \cite{BC92}, \cite[Ch.\ VII]{Be68}, \cite{BG90}, \cite{CFMV03}, \cite{CG06},
\cite{DG92}--\cite{DV95}, \cite{Ge81}, \cite{Ge82}, \cite{Kr49}, \cite{Kr71}, \cite{Le47},
\cite{Lo99}, \cite{Os97}--\cite{Os02}, \cite{Ro90}, \cite{YM01}--\cite{YK78}, and the literature therein, the case of CMV operators with matrix-valued Verblunsky coefficients appears to be a much less explored frontier. The only references we are aware of in this context are Simon's treatise \cite[Part 1, Sect.\ 2.13]{Si04} and a recent preprint by Simon \cite{Si06}.
In the remainder of this paper, ${\mathbb{C}^{m\times m}}$ denotes the space of $m\times m$ matrices with complex-valued entries endowed with the operator norm $\norm{\cdot}_{{\mathbb{C}^{m\times m}}}$ (we use the standard Euclidean norm in ${\mathbb{C}}^m$). The adjoint of an element $\gamma\in{\mathbb{C}^{m\times m}}$ is denoted by $\gamma^*$, and the real and imaginary parts of $\gamma$ are defined as usual by
$\Re(\gamma)=(\gamma+\gamma^*)/2$ and $\Im(\gamma) =(\gamma-\gamma^*)/(2i)$.
\begin{remark} \label{r3.0}
For simplicity of exposition, we find it convenient to use the
following conventions: We denote by $\s{{\mathbb{Z}}}$ the vector space of all
$\mathbb{C}$-valued sequences, and by $\sm{{\mathbb{Z}}}=\s{{\mathbb{Z}}}\otimes\mathbb{C}^m$ the vector
space of all $\mathbb{C}^m$-valued sequences; that is,
\begin{align} \label{3.5A}
\phi=\{\phi(k)\}_{k\in{\mathbb{Z}}}=
\begin{pmatrix}
\vdots\\\phi(-1)\\\phi(0)\\\phi(1)\\\vdots
\end{pmatrix}\in\sm{{\mathbb{Z}}}, \quad
\phi(k)=
\begin{pmatrix}
(\phi(k))_1\\(\phi(k))_2\\\vdots\\(\phi(k))_m
\end{pmatrix}\in\mathbb{C}^m, \; k\in{\mathbb{Z}}.
\end{align}
Moreover, we introduce $\smn{{\mathbb{Z}}}=\sm{{\mathbb{Z}}}\otimes\mathbb{C}^n$, $m,n\in{\mathbb{N}}$, that is,
$\Phi=(\phi_1,\dots,\phi_n)\in\smn{{\mathbb{Z}}}$, where $\phi_j\in\sm{{\mathbb{Z}}}$
for all $j=1,\dots,n$.
We also note that $\smn{{\mathbb{Z}}}=\s{{\mathbb{Z}}}\otimes\mathbb{C}^{m\times n}$, $m,n\in{\mathbb{N}}$;
which is to say that the elements of $\smn{{\mathbb{Z}}}$ can be
identified with the $\mathbb{C}^{m\times n}$-valued sequences,
\begin{align} \label{3.6A}
\Phi=\{\Phi(k)\}_{k\in{\mathbb{Z}}}=
\begin{pmatrix}
\vdots\\\Phi(-1)\\\Phi(0)\\\Phi(1)\\\vdots
\end{pmatrix},
\quad \Phi(k)=
\begin{pmatrix}
(\Phi(k))_{1,1} & \hdots & (\Phi(k))_{1,n}\\
\vdots & & \vdots \\
(\Phi(k))_{m,1} & \hdots & (\Phi(k))_{m,n}
\end{pmatrix}\in\mathbb{C}^{m\times n}, \; k\in{\mathbb{Z}},
\end{align}
by setting $\Phi=(\phi_1,\dots,\phi_n)$, where
\begin{align} \label{3.7A}
\phi_j=
\begin{pmatrix}
\vdots\\\phi_j(-1)\\\phi_j(0)\\\phi_j(1)\\\vdots
\end{pmatrix}\in\sm{{\mathbb{Z}}}, \quad \phi_j(k)=
\begin{pmatrix}
(\Phi(k))_{1,j}\\\vdots\\(\Phi(k))_{m,j}
\end{pmatrix}\in\mathbb{C}^m, \; j=1,\dots,n,\; k\in{\mathbb{Z}}.
\end{align}
For the elements of $\smn{{\mathbb{Z}}}$ we define the right-multiplication by
$n\times n$ matrices with complex-valued entries by
\begin{align} \label{3.3A}
\Phi C = (\phi_1,\dots,\phi_n)
\begin{pmatrix}c_{1,1} & \dots & c_{1,n}\\
\vdots&&\vdots\\c_{n,1} & \dots & c_{n,n}\end{pmatrix} =
\left(\sum_{j=1}^n \phi_j c_{j,1},\dots,\sum_{j=1}^n \phi_j
c_{j,n}\right)\in\smn{{\mathbb{Z}}}
\end{align}
for all $\Phi\in\smn{{\mathbb{Z}}}$ and $C\in\mathbb{C}^{n\times n}$. In addition, for
any linear transformation ${\mathbb{A}}: \sm{{\mathbb{Z}}}\to\sm{{\mathbb{Z}}}$, we define
${\mathbb{A}}\Phi$ for all $\Phi=(\phi_1,\dots,\phi_n)\in\smn{{\mathbb{Z}}}$ by
\begin{align} \label{3.4A}
{\mathbb{A}}\Phi = ({\mathbb{A}}\phi_1,\dots,{\mathbb{A}}\phi_n)\in\smn{{\mathbb{Z}}}.
\end{align}
Given the above conventions, we note the subspace
containment: $\ltm{{\mathbb{Z}}}=\ell^2({\mathbb{Z}})\otimes\mathbb{C}^m\subset\sm{{\mathbb{Z}}}$ and
$\ltmn{{\mathbb{Z}}}=\ell^2({\mathbb{Z}})\otimes\mathbb{C}^{m\times n}\subset\smn{{\mathbb{Z}}}$. We also
note that $\ltm{{\mathbb{Z}}}$ represents a Hilbert space with scalar
product given by
\begin{align} \label{3.2A}
(\phi,\psi)_{\ltm{{\mathbb{Z}}}} = \sum_{k=-\infty}^\infty\sum_{j=1}^m
\overline{(\phi(k))_j}(\psi(k))_j, \quad \phi,\psi\in\ltm{{\mathbb{Z}}}.
\end{align}
Finally, we note that a straightforward modification of the
above definitions also yields the Hilbert space $\ltm{J}$ as well
as the sets $\ltmn{J}$, $\sm{J}$, and $\smn{J}$ for any
$J\subset{\mathbb{Z}}$.
\end{remark}
We start by introducing our basic assumption:
\begin{hypothesis} \label{h3.1}
Let $m\in{\mathbb{N}}$ and assume $\alpha=\{\alpha_k\}_{k\in{\mathbb{Z}}}$ is a sequence of
$m \times m$ matrices with complex entries\footnote{We emphasize that $\alpha_k\in{\mathbb{C}^{m\times m}}$, $k\in{\mathbb{Z}}$, are general (not necessarily normal) matrices.}
and such that
\begin{equation} \label{3.7}
\norm{\alpha_k}_{{\mathbb{C}^{m\times m}}} < 1, \quad k\in{\mathbb{Z}}.
\end{equation}
\end{hypothesis}
Given a sequence $\alpha$ satisfying \eqref{3.7}, we define two sequences
of positive self-adjoint $m\times m$ matrices $\{\rho_k\}_{k\in{\mathbb{Z}}}$
and $\{\widetilde\rho_k\}_{k\in{\mathbb{Z}}}$ by
\begin{align}
\rho_k &= \sqrt{I_m-\alpha_k^*\alpha_k}, \quad k\in{\mathbb{Z}}, \label{3.8}
\\
\widetilde\rho_k &= \sqrt{I_m-\alpha_k\alpha_k^*}, \quad k\in{\mathbb{Z}}, \label{3.9}
\end{align}
and two sequences of $m\times m$ matrices with positive real parts,
$\{a_k\}_{k\in{\mathbb{Z}}}\subset {\mathbb{C}}^{m\times m}$ and
$\{b_k\}_{k\in{\mathbb{Z}}}\subset {\mathbb{C}}^{m\times m}$ by
\begin{align}
a_k &= I_m+\alpha_k, \quad k\in{\mathbb{Z}}, \label{3.10}
\\
b_k &= I_m-\alpha_k, \quad k \in {\mathbb{Z}}. \label{3.11}
\end{align}
Then \eqref{3.7} implies that $\rho_k$ and $\widetilde\rho_k$ are
invertible matrices for all $k\in{\mathbb{Z}}$, and using elementary power series
expansions one verifies the following identities for all $k\in{\mathbb{Z}}$,
\begin{align}
&\widetilde\rho_k^{\pm1}\alpha_k = \alpha_k\rho_k^{\pm1}, \quad
\alpha_k^*\widetilde\rho_k^{\pm1} = \rho_k^{\pm1}\alpha_k^*, \label{3.12}
\\
&a_k^*\widetilde\rho_k^{-2}a_k = a_k\rho_k^{-2}a_k^*, \quad
b_k^*\widetilde\rho_k^{-2}b_k = b_k\rho_k^{-2}b_k^*, \quad
a_k^*\widetilde\rho_k^{-2}b_k + a_k\rho_k^{-2}b_k^* =
b_k^*\widetilde\rho_k^{-2}a_k + b_k\rho_k^{-2}a_k^* = 2I_m. \label{3.13}
\end{align}
According to Simon \cite{Si04}, we call $\alpha_k$ the Verblunsky
coefficients in honor of Verblunsky's pioneering work in the theory
of orthogonal polynomials on the unit circle \cite{Ve35},
\cite{Ve36}.
Next, we introduce a sequence of $2\times 2$ block unitary matrices $\Theta_k$
with $m\times m$ matrix coefficients by
\begin{equation} \label{3.14}
\Theta_k = \begin{pmatrix} -\alpha_k & \widetilde\rho_k \\ \rho_k & \alpha_k^*
\end{pmatrix},
\quad k \in {\mathbb{Z}},
\end{equation}
and two unitary operators ${\mathbb{V}}$ and ${\mathbb{W}}$ on $\ltm{{\mathbb{Z}}}$ by their
matrix representations in the standard basis of $\ltm{{\mathbb{Z}}}$ by
\begin{align} \label{3.15}
{\mathbb{V}} &= \begin{pmatrix} \ddots & & &
\raisebox{-3mm}[0mm][0mm]{\hspace*{-5mm}\Huge $0$} \\ & \Theta_{2k-2} &
& \\ & & \Theta_{2k} & & \\ &
\raisebox{0mm}[0mm][0mm]{\hspace*{-10mm}\Huge $0$} & & \ddots
\end{pmatrix}, \quad
{\mathbb{W}} = \begin{pmatrix} \ddots & & &
\raisebox{-3mm}[0mm][0mm]{\hspace*{-5mm}\Huge $0$}
\\ & \Theta_{2k-1} & & \\ & & \Theta_{2k+1} & & \\ &
\raisebox{0mm}[0mm][0mm]{\hspace*{-10mm}\Huge $0$} & & \ddots
\end{pmatrix},
\end{align}
where
\begin{align}
\begin{pmatrix}
{\mathbb{V}}_{2k-1,2k-1} & {\mathbb{V}}_{2k-1,2k} \\ {\mathbb{V}}_{2k,2k-1} & {\mathbb{V}}_{2k,2k}
\end{pmatrix} = \Theta_{2k},
\quad
\begin{pmatrix}
{\mathbb{W}}_{2k,2k} & {\mathbb{W}}_{2k,2k+1} \\ {\mathbb{W}}_{2k+1,2k} & {\mathbb{W}}_{2k+1,2k+1}
\end{pmatrix} = \Theta_{2k+1},
\quad k\in{\mathbb{Z}}. \label{3.16}
\end{align}
Moreover, we introduce the unitary operator ${\mathbb{U}}$ on
$\ltm{{\mathbb{Z}}}$ as the product of the unitary operators ${\mathbb{V}}$ and ${\mathbb{W}}$ by
\begin{equation} \label{3.17}
{\mathbb{U}} = {\mathbb{V}}{\mathbb{W}},
\end{equation}
or in matrix form in the standard basis of $\ltm{{\mathbb{Z}}}$, by
\begin{align}
{\mathbb{U}} = \begin{pmatrix} \ddots &&\hspace*{-8mm}\ddots
&\hspace*{-10mm}\ddots &\hspace*{-12mm}\ddots &\hspace*{-14mm}\ddots
&&& \raisebox{-3mm}[0mm][0mm]{\hspace*{-6mm}{\Huge $0$}}
\\
&0& -\alpha_{0}\rho_{-1} & -\alpha_{0}\alpha_{-1}^* & -\widetilde\rho_{0}\alpha_{1} &
\widetilde\rho_{0}\widetilde\rho_{1}
\\
&& \rho_{0}\rho_{-1} &\rho_{0}\alpha_{-1}^* & -\alpha_{0}^*\alpha_{1} &
\alpha_{0}^*\widetilde\rho_{1} & 0
\\
&&&0& -\alpha_{2}\rho_{1} & -\alpha_{2}\alpha_{1}^* & -\widetilde\rho_{2}\alpha_{3} &
\widetilde\rho_{2}\widetilde\rho_{3}
\\
&&\raisebox{-4mm}[0mm][0mm]{\hspace*{-6mm}{\Huge $0$}} &&
\rho_{2}\rho_{1} & \rho_{2}\alpha_{1}^* & -\alpha_{2}^*\alpha_{3} &
\alpha_{2}^*\widetilde\rho_{3}&0
\\
&&&&&\hspace*{-14mm}\ddots &\hspace*{-14mm}\ddots
&\hspace*{-14mm}\ddots &\hspace*{-8mm}\ddots &\ddots
\end{pmatrix}. \label{3.18}
\end{align}
Here terms of the form $-\alpha_{2k}\alpha_{2k-1}^*$ and
$-\alpha_{2k}^*\alpha_{2k+1}$, $k\in{\mathbb{Z}}$, represent the diagonal entries
${\mathbb{U}}_{2k-1,2k-1}$ and ${\mathbb{U}}_{2k,2k}$ of the infinite matrix ${\mathbb{U}}$ in
\eqref{3.18}, respectively. We continue to call the operator ${\mathbb{U}}$ on
$\ltm{{\mathbb{Z}}}$ the CMV operator since \eqref{3.14}--\eqref{3.18} in the
context of the scalar-valued semi-infinite (i.e., half-lattice) case
were obtained by Cantero, Moral, and Vel\'azquez in \cite{CMV03} in
2003, but we refer to the discussion in the introduction about the
involved history of these operators.
\begin{lemma} \label{l3.2}
Let $z\in{\mathbb{C}}\backslash\{0\}$ and $\{U(z,k)\}_{k\in{\mathbb{Z}}},
\{V(z,k)\}_{k\in{\mathbb{Z}}}$ be two ${\mathbb{C}^{m\times m}}$-valued sequences. Then the
following items $(i)$--$(iii)$ are equivalent:
\begin{align}
(i) & \quad ({\mathbb{U}} U(z,\cdot))(k) = z U(z,k), \quad ({\mathbb{W}} U(z,\cdot))(k)=z
V(z,k), \quad k\in{\mathbb{Z}}. \label{3.19}
\\
(ii) & \quad ({\mathbb{W}} U(z,\cdot))(k) = z V(z,k), \quad ({\mathbb{V}} V(z,\cdot))(k) =
U(z,k), \quad k\in{\mathbb{Z}}. \label{3.20}
\\
(iii) & \quad \binom{U(z,k)}{V(z,k)} = {\mathbb{T}}(z,k)
\binom{U(z,k-1)}{V(z,k-1)}, \quad k\in{\mathbb{Z}}. \label{3.21}
\end{align}
Here ${\mathbb{U}}$, ${\mathbb{V}}$, and ${\mathbb{W}}$ are understood in the sense of difference
expressions on $\smm{{\mathbb{Z}}}$ rather than difference operators on
$\ltm{{\mathbb{Z}}}$ $($cf.\ Remark \ref{r3.0}$)$ and the transfer matrices
${\mathbb{T}}(z,k)$, $z\in{\C\backslash\{0\}}$, $k\in{\mathbb{Z}}$, are defined by
\begin{equation}
{\mathbb{T}}(z,k) = \begin{cases}
\begin{pmatrix}
\widetilde\rho_{k}^{-1}\alpha_{k} & z\widetilde\rho_{k}^{-1} \\
z^{-1}\rho_{k}^{-1} & \rho_{k}^{-1}\alpha_{k}^*
\end{pmatrix}, & \text{$k$ odd,}
\\
\begin{pmatrix}
\rho_{k}^{-1}\alpha_{k}^* & \rho_{k}^{-1} \\
\widetilde\rho_{k}^{-1} & \widetilde\rho_{k}^{-1}\alpha_{k}
\end{pmatrix}, & \text{$k$ even.}
\end{cases} \label{3.22}
\end{equation}
\end{lemma}
\begin{proof}
The equivalence of \eqref{3.19} and \eqref{3.20} is a consequence of
\eqref{3.17} and equivalence of \eqref{3.20} and \eqref{3.21} is implied by
the following computations:
Assuming $k$ to be odd and utilizing \eqref{3.8}, \eqref{3.9}, and
\eqref{3.12}, one verifies equivalence of the following items
$(i)$--$(v)$:
\begin{align}
(i) & \quad \binom{U(z,k)}{V(z,k)} = {\mathbb{T}}(z,k)
\binom{U(z,k-1)}{V(z,k-1)}.
\\
(ii) & \quad
\begin{cases}
\widetilde\rho_k U(z,k) = \alpha_k U(z,k-1) + zV(z,k-1), \\
\rho_k zV(z,k) = U(z,k-1) + \alpha_k^* zV(z,k-1).
\end{cases}
\\
(iii) & \quad
\begin{cases}
zV(z,k-1) = - \alpha_k U(z,k-1) + \widetilde\rho_k U(z,k), \\
\rho_k zV(z,k) = U(z,k-1) + \alpha_k^*\big(-\alpha_k U(z,k-1) + \widetilde\rho_k
U(z,k)\big).
\end{cases}
\\
(iv) & \quad
\begin{cases}
zV(z,k-1) = - \alpha_k U(z,k-1) + \widetilde\rho_k U(z,k), \\
\rho_k zV(z,k) = \rho_k^2 U(z,k-1) + \rho_k\alpha_k^* U(z,k).
\end{cases}
\\
(v) & \quad z \binom{V(z,k-1)}{V(z,k)} = \Theta_k
\binom{U(z,k-1)}{U(z,k)}.
\intertext{Similarly, assuming $k$ to be even, one verifies that the items
$(vi)$--$(viii)$ are equivalent:}
(vi) & \quad \binom{U(z,k)}{V(z,k)} = {\mathbb{T}}(z,k)
\binom{U(z,k-1)}{V(z,k-1)}.
\\
(vii) & \quad
\begin{cases}
\widetilde\rho_k V(z,k) = \alpha_k V(z,k-1) + U(z,k-1), \\
\rho_k U(z,k) = V(z,k-1) + \alpha_k^* U(z,k-1).
\end{cases}
\\
(viii) & \quad \binom{U(z,k-1)}{U(z,k)} = \Theta_k
\binom{V(z,k-1)}{V(z,k)}.
\end{align}
Finally, taking into account \eqref{3.15} and \eqref{3.16}, one
concludes that
\begin{align}
\Theta_{2k+1} \binom{U(z,2k)}{U(z,2k+1)} = z \binom{V(z,2k)}{V(z,2k+1)},
\quad \Theta_{2k} \binom{V(z,2k-1)}{V(z,2k)} = z
\binom{U(z,2k-1)}{U(z,2k)}, \quad k\in{\mathbb{Z}}
\end{align}
is equivalent to
\begin{align}
({\mathbb{W}} U(z,\cdot))(k) = z V(z,k), \quad ({\mathbb{V}} V(z,\cdot))(k) = U(z,k),
\quad k\in{\mathbb{Z}}.
\end{align}
\end{proof}
We note that in studying solutions of $({\mathbb{U}} U(z,\cdot))(k)=zU(z,k)$
as in Lemma \ref{l3.2}\,$(i)$, the purpose of the additional
relation $({\mathbb{W}} U(z,\cdot))(k)=zV(z,k)$ in \eqref{3.19} is to
introduce a new variable $V$ that improves our understanding of the
structure of such solutions $U$.
If one sets $\alpha_{k_0} = I_m$ for some reference point $k_0\in{\mathbb{Z}}$,
then the operator ${\mathbb{U}}$ splits into a direct sum of two half-lattice
operators ${\mathbb{U}}_{-,k_0-1}$ and ${\mathbb{U}}_{+,k_0}$ acting on
$\ltm{(-\infty,k_0-1]\cap{\mathbb{Z}}}$ and on $\ltm{[k_0,\infty)\cap{\mathbb{Z}}}$,
respectively. Explicitly, one obtains
\begin{align}
{\mathbb{U}} = {\mathbb{U}}_{-,k_0-1} \oplus {\mathbb{U}}_{+,k_0} \, \text{ in } \,
\ltm{(-\infty,k_0-1]\cap{\mathbb{Z}}} \oplus \ltm{[k_0,\infty)\cap{\mathbb{Z}}}. \label{3.33}
\end{align}
(Strictly, speaking, setting $\alpha_{k_0} = I_m$ for some reference
point $k_0\in{\mathbb{Z}}$ contradicts our basic Hypothesis \ref{h3.1}.
However, as long as the exception to Hypothesis \ref{h3.1} refers to
only one site, we will safely ignore this inconsistency in favor of
the notational simplicity it provides by avoiding the introduction of
a properly modified hypothesis on $\{\alpha_k\}_{k\in{\mathbb{Z}}}$.) Similarly,
one obtains ${\mathbb{W}}_{-,k_0-1}$, ${\mathbb{V}}_{-,k_0-1}$ and ${\mathbb{W}}_{+,k_0}$,
${\mathbb{V}}_{+,k_0}$ such that
\begin{equation}
{\mathbb{U}}_{\pm,k_0} = {\mathbb{V}}_{\pm,k_0} {\mathbb{W}}_{\pm,k_0}. \label{3.34}
\end{equation}
\begin{lemma} \label{l3.3}
Let $z\in{\mathbb{C}}\backslash\{0\}$, $k_0\in{\mathbb{Z}}$, and
$\{\widehat P_+(z,k,k_0)\}_{k\geq k_0}$, $\{\widehat R_+(z,k,k_0)\}_{k\geq
k_0}$ be two
${\mathbb{C}^{m\times m}}$-valued sequences. Then the following items $(i)$--$(vi)$ are
equivalent:
\begin{align}
(i)& \quad ({\mathbb{U}}_{+,k_0} \widehat P_+(z,\cdot,k_0))(k) = z \widehat
P_+(z,k,k_0), \quad
({\mathbb{W}}_{+,k_0} \widehat P_+(z,\cdot,k_0))(k) = z \widehat R_+(z,k,k_0), \quad
k\geq k_0. \label{3.35}\\
(ii)& \quad ({\mathbb{W}}_{+,k_0} \widehat P_+(z,\cdot,k_0))(k) = z \widehat
R_+(z,k,k_0), \quad
({\mathbb{V}}_{+,k_0} \widehat R_+(z,\cdot,k_0))(k) = \widehat P_+(z,k,k_0), \quad k\geq
k_0.
\label{3.36}
\\
(iii)& \quad \binom{\widehat P_+(z,k,k_0)}{\widehat R_+(z,k,k_0)} = {\mathbb{T}}(z,k)
\binom{\widehat P_+(z,k-1,k_0)}{\widehat R_+(z,k-1,k_0)}, \quad k > k_0,
\notag \\
& \hspace*{11mm} \text{with initial condition } \,
\widehat P_+(z,k_0,k_0) =
\begin{cases} z \widehat R_+(z,k_0,k_0), & \text{$k_0$ odd}, \\
\widehat R_+(z,k_0,k_0), & \text{$k_0$ even}. \end{cases}\label{3.37}
\end{align}
Next, consider ${\mathbb{C}^{m\times m}}$-valued sequences $\{\widehat P_-(z,k,k_0)\}_{k\leq
k_0}$,
$\{\widehat R_-(z,k,k_0)\}_{k\leq k_0}$. Then the following items
$(iv)$--$(vi)$ are equivalent:
\begin{align}
(iv)& \quad ({\mathbb{U}}_{-,k_0} \widehat P_-(z,\cdot,k_0))(k) = z \widehat
P_-(z,k,k_0), \quad
({\mathbb{W}}_{-,k_0} \widehat P_-(z,\cdot,k_0))(k) = z \widehat R_-(z,k,k_0), \quad
k\leq k_0.
\\
(v)& \quad ({\mathbb{W}}_{-,k_0} \widehat P_-(z,\cdot,k_0))(k) = z \widehat
R_-(z,k,k_0),
\quad ({\mathbb{V}}_{-,k_0} \widehat R_-(z,\cdot,k_0))(k) = \widehat P_-(z,k,k_0), \quad
k\leq
k_0.
\\
(vi)& \quad \binom{\widehat P_-(z,k-1),k_0}{\widehat R_-(z,k-1,k_0)} =
{\mathbb{T}}(z,k)^{-1}
\binom{\widehat P_-(z,k,k_0)}{\widehat R_-(z,k,k_0)}, \quad k \leq k_0, \notag
\\
& \hspace*{11mm} \text{with initial condition } \,
\widehat P_-(z,k_0,k_0) =\begin{cases} - \widehat R_-(z,k_0,k_0), &
\text{$k_0$ odd,}
\\ -z \widehat R_-(z,k_0,k_0), & \text{$k_0$ even.} \end{cases}\label{3.40}
\end{align}
Here ${\mathbb{U}}_{\pm,k_0}$, ${\mathbb{V}}_{\pm,k_0}$, and ${\mathbb{W}}_{\pm,k_0}$ are
understood in the sense of difference expressions on the set
$\smm{{\mathbb{Z}}\cap[k_0,\pm\infty)}$ rather than difference operators on
$\ltm{{\mathbb{Z}}\cap[k_0,\pm\infty)}$ $($cf.\ Remark \ref{r3.0}$)$.
\end{lemma}
\begin{proof}
Equivalence of \eqref{3.35} and \eqref{3.36} is a consequence of
\eqref{3.34}.
Next, repeating the proof of Lemma \ref{l3.2} one obtains that
\begin{align}
({\mathbb{W}}_{+,k_0} \widehat P_+(z,\cdot,k_0))(k) &= z \widehat R_+(z,k,k_0), \quad ({\mathbb{V}}_{+,k_0}
\widehat R_+(z,\cdot,k_0))(k) = \widehat P_+(z,k,k_0), \quad k > k_0, \label{3.41}
\end{align}
is equivalent to
\begin{align}
\binom{\widehat P_+(z,k,k_0)}{\widehat R_+(z,k,k_0)} &= {\mathbb{T}}(z,k)
\binom{\widehat P_+(z,k-1,k_0)}{\widehat R_+(z,k-1,k_0)}, \quad k > k_0. \label{3.42}
\end{align}
Moreover, in the case $k_0$ is odd, the matrices ${\mathbb{V}}_{+,k_0}$ and
${\mathbb{W}}_{+,k_0}$ have the structure,
\begin{align}
{\mathbb{V}}_{+,k_0} =
\begin{pmatrix}
\Theta_{k_0+1} & & \raisebox{-1mm}[0mm][0mm]{\hspace*{-2mm}{\huge $0$}}
\\ & \Theta_{k_0+3} & \\
\raisebox{-1mm}[0mm][0mm]{\hspace*{0mm}{\huge $0$}} && \ddots
\end{pmatrix},
\quad {\mathbb{W}}_{+,k_0} =
\begin{pmatrix}
I_m & & \raisebox{-1mm}[0mm][0mm]{\hspace*{-2mm}{\huge $0$}}
\\ & \Theta_{k_0+2} & \\ &
\raisebox{-1mm}[0mm][0mm]{\hspace*{-14mm}{\huge $0$}} &\ddots
\end{pmatrix},
\end{align}
and hence,
\begin{equation}
({\mathbb{W}}_{+,k_0} \widehat P_+(z,\cdot,k_0))(k_0) = z \widehat R_+(z,k_0,k_0) \label{3.44}
\end{equation}
is equivalent to
\begin{equation}
\widehat P_+(z,k_0,k_0) = z \widehat R_+(z,k_0,k_0). \label{3.45}
\end{equation}
In the case $k_0$ is even, the matrices $V_{+,k_0}$ and $W_{+,k_0}$
have the structure,
\begin{align}
{\mathbb{V}}_{+,k_0} =
\begin{pmatrix}
I_m & & \raisebox{-1mm}[0mm][0mm]{\hspace*{-2mm}{\huge $0$}}
\\ & \Theta_{k_0+2} & \\ &
\raisebox{-1mm}[0mm][0mm]{\hspace*{-14mm}{\huge $0$}} &\ddots
\end{pmatrix},
\quad {\mathbb{W}}_{+,k_0} =
\begin{pmatrix}
\Theta_{k_0+1} & & \raisebox{-1mm}[0mm][0mm]{\hspace*{-2mm}{\huge $0$}}
\\ &
\Theta_{k_0+3} & \\
\raisebox{-1mm}[0mm][0mm]{\hspace*{0mm}{\huge $0$}} & & \ddots
\end{pmatrix},
\end{align}
and hence,
\begin{equation}
({\mathbb{V}}_{+,k_0} \widehat R_+(z,\cdot,k_0))(k_0) = \widehat P_+(z,k_0,k_0) \label{3.47}
\end{equation}
is equivalent to
\begin{equation}
\widehat P_+(z,k_0,k_0) = \widehat R_+(z,k_0,k_0). \label{3.48}
\end{equation}
Thus, one infers the equivalence of \eqref{3.36} and \eqref{3.37}
from the equivalence of \eqref{3.41} and \eqref{3.42} with
\eqref{3.44}--\eqref{3.45} and \eqref{3.47}--\eqref{3.48}.
The results for $\{\widehat P_-(z,k,k_0)\}_{k\leq k_0}$ and
$\{\widehat R_-(z,k,k_0)\}_{k\leq k_0}$ are proved analogously.
\end{proof}
Analogous comments to those made right after the proof of Lemma
\ref{l3.2} apply in the present context of Lemma \ref{l3.3}.
Next, we denote by
$\Big(\begin{smallmatrix}P_\pm(z,k,k_0) \\
R_\pm(z,k,k_0)\end{smallmatrix}\Big)_{k\in{\mathbb{Z}}}$ and
$\Big(\begin{smallmatrix} Q_\pm(z,k,k_0) \\ S_\pm(z,k,k_0)
\end{smallmatrix}\Big)_{k\in{\mathbb{Z}}}$,
$z\in{\mathbb{C}}\backslash\{0\}$, four linearly independent solutions of
\eqref{3.21} satisfying the following initial conditions:
\begin{align}
\binom{P_+(z,k_0,k_0)}{R_+(z,k_0,k_0)} &=
\begin{cases}
\binom{zI_m}{I_m}, & \text{$k_0$ odd,} \\[1mm]
\binom{I_m}{I_m}, & \text{$k_0$ even,}
\end{cases} \quad\;\;
\binom{Q_+(z,k_0,k_0)}{S_+(z,k_0,k_0)} =
\begin{cases}
\binom{zI_m}{-I_m}, & \text{$k_0$ odd,} \\[1mm]
\binom{-I_m}{I_m}, & \text{$k_0$ even.}
\end{cases} \label{3.49}
\\
\binom{P_-(z,k_0,k_0)}{R_-(z,k_0,k_0)} &=
\begin{cases}
\binom{I_m}{-I_m}, & \text{$k_0$ odd,} \\[1mm]
\binom{-zI_m}{I_m}, & \text{$k_0$ even,}
\end{cases} \quad
\binom{Q_-(z,k_0,k_0)}{S_-(z,k_0,k_0)} =
\begin{cases}
\binom{I_m}{I_m}, & \text{$k_0$ odd, }\\[1mm]
\binom{zI_m}{I_m}, & \text{$k_0$ even.}
\end{cases} \label{3.50}
\end{align}
Then $P_\pm(z,k,k_0)$, $Q_\pm(z,k,k_0)$,
$R_\pm(z,k,k_0)$, and $S_\pm(z,k,k_0)$, $k,k_0\in{\mathbb{Z}}$, are
${\mathbb{C}^{m\times m}}$-valued Laurent polynomials in $z$. In particular, one computes
\begin{align}\label{Table}
\begin{array}{|c|c|c|c|}
\hline k & k_0-1 & k_0 \text{ odd} & k_0+1
\\ \hline \ph{\Bigg|}
\displaystyle \binom{P_+(z,k,k_0)}{R_+(z,k,k_0)} & \displaystyle
\binom{z\rho_{k_0}^{-1}(I_m-\alpha_{k_0}^*)}
{\widetilde\rho_{k_0}^{-1}(I_m-\alpha_{k_0})} & \displaystyle
\binom{zI_m}{I_m} & \displaystyle
\binom{\rho_{k_0+1}^{-1}(I_m+z\alpha_{k_0+1}^*)}
{\widetilde\rho_{k_0+1}^{-1}(zI_m+\alpha_{k_0+1})}
\\ \hline \ph{\Bigg|}
\displaystyle \binom{Q_+(z,k,k_0)}{S_+(z,k,k_0)} & \displaystyle
\binom{z\rho_{k_0}^{-1}(-I_m-\alpha_{k_0}^*)}
{\widetilde\rho_{k_0}^{-1}(I_m+\alpha_{k_0})} & \displaystyle
\binom{zI_m}{-I_m} & \displaystyle
\binom{\rho_{k_0+1}^{-1}(-I_m+z\alpha_{k_0+1}^*)}
{\widetilde\rho_{k_0+1}^{-1}(zI_m-\alpha_{k_0+1})}
\\ \hline \ph{\Bigg|}
\displaystyle \binom{P_-(z,k,k_0)}{R_-(z,k,k_0)} & \displaystyle
\binom{\rho_{k_0}^{-1}(-zI_m-\alpha_{k_0}^*)}
{\widetilde\rho_{k_0}^{-1}(\frac1zI_m+\alpha_{k_0})} & \displaystyle
\binom{I_m}{-I_m} & \displaystyle
\binom{\rho_{k_0+1}^{-1}(-I_m+\alpha_{k_0+1}^*)}
{\widetilde\rho_{k_0+1}^{-1}(I_m-\alpha_{k_0+1})}
\\ \hline \ph{\Bigg|}
\displaystyle \binom{Q_-(z,k,k_0)}{S_-(z,k,k_0)} & \displaystyle
\binom{\rho_{k_0}^{-1}(zI_m-\alpha_{k_0}^*)}
{\widetilde\rho_{k_0}^{-1}(\frac1zI_m-\alpha_{k_0})} & \displaystyle
\binom{I_m}{I_m} & \displaystyle
\binom{\rho_{k_0+1}^{-1}(I_m+\alpha_{k_0+1}^*)}
{\widetilde\rho_{k_0+1}^{-1}(I_m+\alpha_{k_0+1})}
\\ \hline \hline
k & k_0-1 & k_0 \text{ even} & k_0+1
\\ \hline \ph{\Bigg|}
\displaystyle \binom{P_+(z,k,k_0)}{R_+(z,k,k_0)} & \displaystyle
\binom{\widetilde\rho_{k_0}^{-1}(I_m-\alpha_{k_0})}{\rho_{k_0}^{-1}(I_m-\alpha_{k_0}^*)}
& \displaystyle \binom{I_m}{I_m} & \displaystyle
\binom{\widetilde\rho_{k_0+1}^{-1}(zI_m+\alpha_{k_0+1})}
{\rho_{k_0+1}^{-1}(\frac1zI_m+\alpha_{k_0+1}^*)}
\\ \hline \ph{\Bigg|}
\displaystyle \binom{Q_+(z,k,k_0)}{S_+(z,k,k_0)} & \displaystyle
\binom{\widetilde\rho_{k_0}^{-1}(I_m+\alpha_{k_0})}
{\rho_{k_0}^{-1}(-I_m-\alpha_{k_0}^*)} & \displaystyle \binom{-I_m}{I_m}
& \displaystyle \binom{\widetilde\rho_{k_0+1}^{-1}(zI_m-\alpha_{k_0+1})}
{\rho_{k_0+1}^{-1}(-\frac1zI_m+\alpha_{k_0+1}^*)}
\\ \hline \ph{\Bigg|}
\displaystyle \binom{P_-(z,k,k_0)}{R_-(z,k,k_0)} & \displaystyle
\binom{\widetilde\rho_{k_0}^{-1}(I_m+z\alpha_{k_0})}
{\rho_{k_0}^{-1}(-zI_m-\alpha_{k_0}^*)} & \displaystyle
\binom{-zI_m}{I_m} & \displaystyle
\binom{z\widetilde\rho_{k_0+1}^{-1}(I_m-\alpha_{k_0+1})}
{\rho_{k_0+1}^{-1}(-I_m+\alpha_{k_0+1}^*)}
\\ \hline \ph{\Bigg|}
\displaystyle \binom{Q_-(z,k,k_0)}{S_-(z,k,k_0)} & \displaystyle
\binom{\widetilde\rho_{k_0}^{-1}(I_m-z\alpha_{k_0})}
{\rho_{k_0}^{-1}(zI_m-\alpha_{k_0}^*)} & \displaystyle \binom{zI_m}{I_m}
& \displaystyle \binom{z\widetilde\rho_{k_0+1}^{-1}(I_m+\alpha_{k_0+1})}
{\rho_{k_0+1}^{-1}(I_m+\alpha_{k_0+1}^*)}
\\ \hline
\end{array}
\end{align}
\begin{remark} \label{r2.4}
Subsequently, we will have to refer to the leading-order terms of certain matrix-valued Laurent polynomials at various occasions. To put this in precise terms we now introduce the following conventions: We will refer to the terms
\begin{align}
\begin{cases}
z^{-(k+1)/2}, & k \text{ odd},
\\[1mm]
z^{k/2}, & k \text{ even},
\end{cases}
\end{align}
as the leading-order terms of the Laurent polynomials
\begin{align}
\begin{cases}
z^{-1}P_+(z,k_0+k,k_0),\; R_-(z,k_0-k,k_0),
\\
z^{-1}Q_+(z,k_0+k,k_0),\; S_-(z,k_0-k,k_0), & k_0 \text{ odd},
\\
R_+(z,k_0+k,k_0), \; z^{-1}P_-(z,k_0-k,k_0),
\\
S_+(z,k_0+k,k_0), \; z^{-1}Q_-(z,k_0-k,k_0), & k_0 \text{ even},
\end{cases}
\end{align}
and similarly, we will refer to the terms
\begin{align}
\begin{cases}
z^{(k+1)/2}, & k \text{ odd},
\\[1mm]
z^{-k/2}, & k \text{ even},
\end{cases}
\end{align}
as the leading-order term of the Laurent polynomials
\begin{align}
\begin{cases}
R_+(z,k_0+k,k_0), \; P_-(z,k_0-k,k_0),
\\
S_+(z,k_0+k,k_0), \; Q_-(z,k_0-k,k_0), & k_0 \text{ odd},
\\
P_+(z,k_0+k,k_0), \; R_-(z,k_0-k,k_0),
\\
Q_+(z,k_0+k,k_0), \; S_-(z,k_0-k,k_0), & k_0 \text{ even}.
\end{cases}
\end{align}
\end{remark}
\begin{remark}
We note that Lemmas \ref{l3.2} and \ref{l3.3} are crucial for many
of the proofs to follow. For instance, we note that the equivalence
of items $(i)$ and $(iii)$ in Lemma \ref{l3.2} proves that for each
$z\in{\mathbb{C}}\backslash\{0\}$, any solutions $\{U(z,k)\}_{k\in{\mathbb{Z}}}$ of
${\mathbb{U}} U(z,\cdot)=zU(z,\cdot)$ can be expressed as a linear
combinations of $P_{+}(z,\cdot,k_0)$ and $Q_{+}(z,\cdot,k_0)$ (or
$P_{-}(z,\cdot,k_0)$ and $Q_{-}(z,\cdot,k_0)$) with $z$-dependent
right-multiple ${\mathbb{C}^{m\times m}}$-valued coefficients. This equivalence also
proves that any solution of ${\mathbb{U}} U(z,\cdot)=zU(z,\cdot)$ is
determined by the values of $U$ and the auxiliary variable $V$ at a
site $k_0$. In the context of Lemma \ref{l3.3}, we remark that its
importance lies in the fact that it shows that in the case of
half-lattice CMV operators, the analogous equations have solutions,
which up to right-multiplication by $z$-dependent ${\mathbb{C}^{m\times m}}$-valued
coefficients, are given by $\{P_{\pm}(z,k,k_0)\}_{k\in{\mathbb{Z}}}$ for
each $z\in{\mathbb{C}}\backslash\{0\}$.\ Consequently, the corresponding
solutions are determined by their value at a single site $k_0$.
\end{remark}
Next, we introduce the modified matrix-valued Laurent polynomials
$\widetilde P_\pm(z,k,k_0)$ and $\widetilde Q_\pm(z,k,k_0)$, $z\in{\C\backslash\{0\}}$, $k,k_0\in{\mathbb{Z}}$,
by
\begin{align}
\widetilde P_+(z,k,k_0) &=
\begin{cases}
P_+(z,k,k_0)/z, & \text{$k_0$ odd,} \\
P_+(z,k,k_0), & \text{$k_0$ even,}
\end{cases}
\quad \widetilde P_-(z,k,k_0) =
\begin{cases}
P_-(z,k,k_0), & \text{$k_0$ odd,} \\
-P_-(z,k,k_0)/z, & \text{$k_0$ even,}
\end{cases} \label{3.49a}
\\
\widetilde Q_+(z,k,k_0) &=
\begin{cases}
Q_+(z,k,k_0)/z, & \text{$k_0$ odd,} \\
Q_+(z,k,k_0), & \text{$k_0$ even,}
\end{cases}
\quad \widetilde Q_-(z,k,k_0) =
\begin{cases}
Q_-(z,k,k_0), & \text{$k_0$ odd,} \\
-Q_-(z,k,k_0)/z, & \text{$k_0$ even.}
\end{cases} \label{3.50a}
\end{align}
In the remainder of this paper we use the following abbreviations for subarcs $A_{\zeta}$ of ${\partial\hspace*{.2mm}\mathbb{D}}$,
\begin{equation}
A_{\zeta}=\big\{e^{i\phi}\in{\partial\hspace*{.2mm}\mathbb{D}}\,\big|\, 0\leq \phi\leq\theta\big\}, \quad \zeta=e^{i\theta},
\; \theta \in [0,2\pi).
\end{equation}
The next auxiliary result is of importance in proving orthonormality of the matrix-valued Laurent polynomials $P_\pm$ and $R_\pm$.
\begin{lemma} \label{l3.5}
Let $\{F_\pm(\cdot,k,k_0)\}_{k\gtreqless k_0}$ denote two sequences
of ${\mathbb{C}^{m\times m}}$-valued functions of bounded variation with $F_\pm(1,k,k_0)=0$
for all $k\gtreqless k_0$ that satisfy
\begin{align}
({\mathbb{U}}_{\pm,k_0} F_\pm(\zeta,\cdot,k_0))(k) = \int_{A_{\zeta}}
dF_\pm(\zeta',k,k_0) \, \zeta', \quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\; k\gtreqless k_0,
\end{align}
where ${\mathbb{U}}_{\pm,k_0}$ are understood in the sense of difference
expressions on $\smm{{\mathbb{Z}}\cap[k_0,\pm\infty)}$ rather than difference
operators on $\ltm{{\mathbb{Z}}\cap[k_0,\pm\infty)}$ $($cf.\ Remark
\ref{r3.0}$)$. Then, $F_\pm(\cdot,k,k_0)$ also satisfy
\begin{align}
F_\pm(\zeta,k,k_0) = \int_{A_{\zeta}} \widetilde
P_\pm(\zeta',k,k_0)\,dF_\pm(\zeta',k_0,k_0), \quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\;
k\gtreqless k_0.
\end{align}
\end{lemma}
\begin{proof}
Let $\{G_\pm(\cdot,k,k_0)\}_{k\gtreqless k_0}$ denote the
two sequences of ${\mathbb{C}^{m\times m}}$-valued functions,
\begin{align}
G_\pm(\zeta,k,k_0) = \int_{A_{\zeta}} \widetilde
P_\pm(\zeta',k,k_0)dF_\pm(\zeta',k_0,k_0), \quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\;
k\gtreqless k_0.
\end{align}
Then it suffices to prove that $F_\pm(\zeta,k,k_0) =
G_\pm(\zeta,k,k_0)$, $\zeta\in{\partial\hspace*{.2mm}\mathbb{D}}$, $k\gtreqless k_0$.
First, we note that according to \eqref{3.49}, \eqref{3.50}, and
\eqref{3.49a}, $\widetilde P_\pm(\zeta,k_0,k_0) = I_m$, and hence,
\begin{align}
G_\pm(\zeta,k_0,k_0) = \int_{A_{\zeta}} dF_\pm(\zeta',k_0,k_0) =
F_\pm(\zeta,k_0,k_0), \quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}}.
\end{align}
Moreover,
\begin{align}
({\mathbb{U}}_{\pm,k_0} G_\pm(\zeta,\cdot,k_0))(k) &= \int_{A_{\zeta}}
({\mathbb{U}}_{\pm,k_0}\widetilde P_\pm(\zeta',\cdot,k_0))(k)dF_\pm(\zeta',k_0,k_0) \notag
\\
&= \int_{A_{\zeta}} dG_\pm(\zeta',k,k_0) \, \zeta', \quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\;
k\gtreqless k_0.
\end{align}
Next, defining $K_\pm(\zeta,k,k_0) =
F_\pm(\zeta,k,k_0)-G_\pm(\zeta,k,k_0)$, $\zeta\in{\partial\hspace*{.2mm}\mathbb{D}}$, $k\gtreqless
k_0$, one obtains
\begin{align*}
K_\pm(\zeta,k_0,k_0) = 0 \quad \text{and}\quad ({\mathbb{U}}_{\pm,k_0}
K_\pm(\zeta,\cdot,k_0))(k) = \int_{A_{\zeta}} dK_\pm(\zeta',k,k_0) \, \zeta' , \quad
\zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\; k\gtreqless k_0,
\end{align*}
or equivalently,
\begin{align}
K_\pm(\zeta,k_0,k_0) = 0 \quad \text{and}\quad ({\mathbb{U}}_{\pm,k_0}
K_\pm(\zeta,\cdot,k_0))(k) = (\L \, K_\pm(\cdot,k,k_0))(\zeta), \quad
\zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\; k\gtreqless k_0, \label{3.53}
\end{align}
where $\L$ denotes the boundedly invertible operator on ${\mathbb{C}^{m\times m}}$-valued
functions $K$ of bounded variation defined by
\begin{align}
(\L \, K)(\zeta) = \int_{A_{\zeta}} dK(\zeta')\, \zeta', \quad (\L^{-1} K)(\zeta) =
\int_{A_{\zeta}} dK(\zeta') \, {\zeta'}^{-1}.
\end{align}
Finally, since, $\L$ commutes with all constant $m\times m$ matrices,
one can repeat the proof of Lemma \ref{l3.3} with $z$ replaced by
$\L$ and obtain that \eqref{3.53} has the unique solution
$K_\pm(\zeta,k,k_0)=0$, $\zeta\in{\partial\hspace*{.2mm}\mathbb{D}}$, $k\gtreqless k_0$, and hence,
$F_\pm(\zeta,k,k_0) = G_\pm(\zeta,k,k_0)$, $\zeta\in{\partial\hspace*{.2mm}\mathbb{D}}$, $k\gtreqless
k_0$.
\end{proof}
Next, following \cite{Be68} (see also \cite{BG90}), we prove a matrix-valued version
of the ``orthogonality" relation for matrix-valued Laurent polynomials
$P_\pm$ and $R_\pm$.
Let $\Delta_k=\{\Delta_k(\ell)\}_{\ell\in{\mathbb{Z}}}\in\smm{{\mathbb{Z}}}$, $k\in{\mathbb{Z}}$, denote
the sequences of $m\times m$ matrices defined by
\begin{align}
(\Delta_k)(\ell) =
\begin{cases}
I_m, & \ell=k,\\
0, & \ell\neq k,
\end{cases}
\quad k,\ell\in{\mathbb{Z}}.
\end{align}
Then using right-multiplication by $m\times m$ matrices on
$\smm{{\mathbb{Z}}}$ defined in Remark \ref{r3.0}, we get the
identity
\begin{align} \label{B.1a}
(\Delta_k X)(\ell) =
\begin{cases}
X, & \ell=k,\\
0, & \ell\neq k,
\end{cases} \quad X\in{\mathbb{C}^{m\times m}},
\end{align}
and hence consider $\Delta_k$ as a map $\Delta_k\colon {\mathbb{C}^{m\times m}}\to\smm{{\mathbb{Z}}}$. In
addition, we introduce the map $\Delta_k^*\colon \smm{{\mathbb{Z}}}\to{\mathbb{C}^{m\times m}}$,
$k\in{\mathbb{Z}}$, defined by
\begin{align} \label{B.1b}
\Delta_k^* \Phi = \Phi(k), \,\text{ where }\,
\Phi=\{\Phi(k)\}_{k\in{\mathbb{Z}}}\in\smm{{\mathbb{Z}}}.
\end{align}
Similarly, one introduces the corresponding maps with ${\mathbb{Z}}$ replaced by
$[k_0,\pm\infty)\cap{\mathbb{Z}}$, $k_0\in{\mathbb{Z}}$, which, for notational brevity, we
will also denote by $\Delta_k$ and $\Delta_k^*$, respectively.
Next, we call sequences of $C^{m\times m}$-valued
functions $\{\Phi_\pm(\cdot,k,k_0)\}_{k\gtreqless k_0}$
orthonormal\footnote{This is denoted by pseudo-orthonormality in
\cite[Sect.~VII.2.6]{Be68}} on ${\partial\hspace*{.2mm}\mathbb{D}}$ with respect to some
${\mathbb{C}^{m\times m}}$-valued measures $d\Omega_\pm(\cdot,k_0)$, defined on ${\partial\hspace*{.2mm}\mathbb{D}}$, if
the following identity holds for all $k,k'\gtreqless k_0$,
\begin{align}
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}
\Phi_\pm(\zeta,k,k_0)\,d\Omega_\pm(\zeta,k_0)\,\Phi_\pm(\zeta,k',k_0)^*
&= \delta_{k,k'}I_m.
\end{align}
We will also call the sequences of ${\mathbb{C}^{m\times m}}$-valued functions
$\{\Phi_\pm(\cdot,k,k_0)\}_{k\gtreqless k_0}$ complete with respect
to the measures $d\Omega_\pm(\cdot,k_0)$ if the collections
of $\mathbb{C}^m$-valued functions
\begin{align}
\left\{\phi_{\pm,j}(\cdot,k,k_0)=
\begin{pmatrix}
(\Phi_\pm(\cdot,k,k_0))_{1,j}\\\vdots\\(\Phi_\pm(\cdot,k,k_0))_{m,j}
\end{pmatrix}\right\}_{ j=1,\dots,m, \; k\gtreqless k_0}
\end{align}
form complete systems in $\Ltm{\pm}$.
\begin{lemma} \label{l3.6}
Let $k_0\in{\mathbb{Z}}$. The sets of ${\mathbb{C}^{m\times m}}$-valued Laurent polynomials
$\{P_\pm(\cdot,k,k_0)^*\}_{k\gtreqless k_0}$ and
$\{R_\pm(\cdot,k,k_0)^*\}_{k\gtreqless k_0}$ form complete
orthonormal systems on ${\partial\hspace*{.2mm}\mathbb{D}}$ with respect to ${\mathbb{C}^{m\times m}}$-valued measures
$d\Omega_\pm(\cdot,k_0)$ defined by
\begin{equation}
d\Omega_\pm(\zeta,k_0) = d(\Delta_{k_0}^* E_{{\mathbb{U}}_{\pm,k_0}}(\zeta)\Delta_{k_0}),
\quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}}, \label{3.54}
\end{equation}
where $E_{{\mathbb{U}}_{\pm,k_0}}(\cdot)$ denotes the family of spectral projections of
the half-lattice unitary operators ${\mathbb{U}}_{\pm,k_0}$,
\begin{equation}
{\mathbb{U}}_{\pm,k_0}=\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}dE_{{\mathbb{U}}_{\pm,k_0}}(\zeta)\,\zeta. \label{3.55}
\end{equation}
Explicitly, $P_\pm$ and $R_\pm$ satisfy,
\begin{align}
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}
P_\pm(\zeta,k,k_0)\,d\Omega_\pm(\zeta,k_0)\,P_\pm(\zeta,k',k_0)^* &=
\delta_{k,k'}I_m, \quad k,k' \gtreqless k_0, \label{3.56}
\\
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}
R_\pm(\zeta,k,k_0)\,d\Omega_\pm(\zeta,k_0)\,R_\pm(\zeta,k',k_0)^* &=
\delta_{k,k'}I_m, \quad k,k' \gtreqless k_0. \label{3.57}
\end{align}
\end{lemma}
\begin{proof}
Fix an integer $k_1\gtreqless k_0$ and let
$\{F_\pm(\cdot,k,k_0)\}_{k\gtreqless k_0}$ denote two ${\mathbb{C}^{m\times m}}$-valued
sequences of functions of bounded variation,
\begin{align}
F_\pm(\zeta,k,k_0) = \Delta_{k}^* E_{{\mathbb{U}}_{\pm,k_0}}(\zeta)\Delta_{k_1}, \quad
\zeta\in{\partial\hspace*{.2mm}\mathbb{D}}, \; k\gtreqless k_0.
\end{align}
Then,
\begin{align}
({\mathbb{U}}_{\pm,k_0} F_\pm(\zeta,\cdot,k_0))(k) &=
({\mathbb{U}}_{\pm,k_0}E_{{\mathbb{U}}_{\pm,k_0}}(\zeta)\Delta_{k_1})(k) = \left(\int_{A_{\zeta}}
dE_{{\mathbb{U}}_{\pm,k_0}}(\zeta')\, \zeta' \Delta_{k_1}\right)(k)
\\ &=
\int_{A_{\zeta}} d\big(\Delta_k^* E_{{\mathbb{U}}_{\pm,k_0}}(\zeta')\Delta_{k_1}\big) \, \zeta' =
\int_{A_{\zeta}} dF_\pm(\zeta',k,k_0) \, \zeta', \quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\;
k\gtreqless k_0, \notag
\end{align}
and hence it follows from Lemma \ref{l3.5} that
\begin{align}
F_\pm(\zeta,k,k_0) = \int_{A_{\zeta}} \widetilde
P_\pm(\zeta',k,k_0)\,dF_\pm(\zeta',k_0,k_0), \quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}},
\; k\gtreqless k_0,
\end{align}
or equivalently,
\begin{align}
\Delta_{k}^* E_{{\mathbb{U}}_{\pm,k_0}}(\zeta)\Delta_{k_1} = \int_{A_{\zeta}} \widetilde
P_\pm(\zeta',k,k_0)\,d\big(\Delta_{k_0}^*
E_{{\mathbb{U}}_{\pm,k_0}}(\zeta')\Delta_{k_1}\big), \quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\;
k\gtreqless k_0.
\end{align}
In particular, taking $k_1=k'$ and $k_1=k_0$, one obtains,
respectively,
\begin{align}
\Delta_{k}^* E_{{\mathbb{U}}_{\pm,k_0}}(\zeta)\Delta_{k'} &= \int_{A_{\zeta}} \widetilde
P_\pm(\zeta',k,k_0)\,d\big(\Delta_{k_0}^*
E_{{\mathbb{U}}_{\pm,k_0}}(\zeta')\Delta_{k'}\big), \quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\;
k\gtreqless k_0, \label{3.62}
\end{align}
and
\begin{align}
\Delta_{k'}^* E_{{\mathbb{U}}_{\pm,k_0}}(\zeta)\Delta_{k_0} &= \int_{A_{\zeta}} \widetilde
P_\pm(\zeta',k',k_0)\,d\big(\Delta_{k_0}^*
E_{{\mathbb{U}}_{\pm,k_0}}(\zeta')\Delta_{k_0}\big) \notag
\\ &=
\int_{A_{\zeta}} \widetilde P_\pm(\zeta',k',k_0)\,d\Omega_\pm(\zeta',k_0), \quad
\zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\; k' \gtreqless k_0. \label{3.64}
\end{align}
Taking adjoints in \eqref{3.64} one also obtains
\begin{align}
\Delta_{k_0}^* E_{{\mathbb{U}}_{\pm,k_0}}(\zeta)\Delta_{k'} &= \int_{A_{\zeta}}
d\Omega_\pm(\zeta',k_0)\,\widetilde P_\pm(\zeta',k',k_0)^*, \quad
\zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\; k' \gtreqless k_0. \label{3.65}
\end{align}
Thus, inserting \eqref{3.49a} and \eqref{3.65} into \eqref{3.62} and
letting $\theta\to2\pi$, $\zeta=e^{i\theta}$, yields \eqref{3.56},
\begin{align}
\delta_{k,k'} I_m &= \oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \widetilde
P_\pm(\zeta,k,k_0)\,d\Omega_\pm(\zeta,k_0)\,\widetilde P_\pm(\zeta,k',k_0)^*
\notag
\\
&= \oint_{{\partial\hspace*{.2mm}\mathbb{D}}}
P_\pm(\zeta,k,k_0)\,d\Omega_\pm(\zeta,k_0)\,P_\pm(\zeta,k',k_0)^*, \quad
k,k'\gtreqless k_0.
\end{align}
Finally, \eqref{3.57} is a consequence of \eqref{3.56} and the
relation
\begin{align}
R_\pm(z,k,k_0) = \frac1z ({\mathbb{W}}_{\pm,k_0}P_\pm(z,\cdot,k_0))(k), \quad
z\in{\C\backslash\{0\}},\; k\gtreqless k_0,
\end{align}
where ${\mathbb{W}}_{\pm,k_0}$ are the unitary block diagonal semi-infinite
matrices defined in \eqref{3.34}.
To prove completeness of $\{P_\pm(\cdot,k,k_0)^*\}_{k\gtreqless
k_0}$ and $\{R_\pm(\cdot,k,k_0)^*\}_{k\gtreqless k_0}$ we first note
the subsequent fact that can be inferred from the definitions of $P_\pm$ and
$R_\pm$ and, in particular, from \eqref{3.21}, \eqref{3.22}, \eqref{3.49},
and \eqref{3.50},
\begin{align}
{\text{\rm span}}\{P_\pm(\zeta,k,k_0)^*\}_{k\gtreqless k_0} &=
{\text{\rm span}}\{R_\pm(\zeta,k,k_0)^*\}_{k\gtreqless k_0} = {\text{\rm span}}\left\{\zeta^k
I_m\right\}_{k\in{\mathbb{Z}}}.
\end{align}
Hence, it suffices to prove that $\big\{\zeta^k I_m\big\}_{k\in{\mathbb{Z}}}$
are complete with respect to $d\Omega_\pm(\cdot,k_0)$. Suppose
$F\in\Ltm{\pm}$ is orthogonal to all columns of $\zeta^kI_m$ for all
$k\in{\mathbb{Z}}$, that is,
\begin{equation}
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \zeta^{-k}
d\Omega(\zeta,k_0)\,F(\zeta)=
\begin{pmatrix}0\\\vdots\\0\end{pmatrix}\in\mathbb{C}^m,
\quad k\in{\mathbb{Z}}.
\end{equation}
Note that for a scalar complex-valued measure $d\omega$ equalities
$\oint d\omega(\zeta)\,\zeta^n=0$, $n\in{\mathbb{Z}}$, imply that $\oint
d\Re(\omega(\zeta))\,\zeta^n=\oint d\Im(\omega(\zeta))\,\zeta^n = 0$, and hence one
concludes from \cite[p.\ 24]{Du83}) that $d\omega=0$. Applying this
argument to $d(\Omega_{\pm}(\cdot,k_0)F(\cdot))_\ell$,
$\ell=1,\dots,m$, one obtains
\begin{align}
d\Omega_{\pm}(\cdot,k_0)F(\cdot) &=
\begin{pmatrix}0\\\vdots\\0\end{pmatrix}\in\mathbb{C}^m.
\end{align}
Multiplying by $F(\cdot)^*$ on the left and integrating over the
unit circle then yields
\begin{align}
\|F\|_{\Ltm{\pm}}^2=\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} F(\zeta)^* \,d\Omega_\pm(\zeta,k_0)\,
F(\zeta) = 0.
\end{align}
\end{proof}
We note that $d\Omega_\pm(\cdot,k_0)$, $k_0\in{\mathbb{Z}}$, defined in
\eqref{3.54} are normalized, nonnegative, nondegenerate,
${\mathbb{C}^{m\times m}}$-valued measures supported on infinite subsets of ${\partial\hspace*{.2mm}\mathbb{D}}$,
that is, for any ${\mathbb{C}^{m\times m}}$-valued Laurent polynomial $P(z)$ the
following properties hold,
\begin{align}
(i)&\; \oint_{{\partial\hspace*{.2mm}\mathbb{D}}} d\Omega_\pm(\zeta,k_0) = I_m \;\text{ and }\;
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} P(\zeta)\,d\Omega_\pm(\zeta,k_0)\,P(\zeta)^* \geq 0.
\\
(ii)&\;\text{ If $P(z)= z^{-n}A_{-n}+...+z^n A_n$ and either $A_n$ or $A_{-n}$ is invertible,} \\
&\; \text{ then } \, \oint_{{\partial\hspace*{.2mm}\mathbb{D}}} P(\zeta)\,d\Omega_\pm(\zeta,k_0)\,P(\zeta)^* > 0.
\\
(iii)&\;\text{ If }\, \oint_{{\partial\hspace*{.2mm}\mathbb{D}}}
P(\zeta)\,d\Omega_\pm(\zeta,k_0)\,P(\zeta)^* = 0 \, \text{ then }\,
P(z)=0.
\end{align}
The infinite support property of the spectral measure is a consequence of
the fact that we have infinitely many linearly independent
orthogonal Laurent polynomials $P_\pm$. Property $(i)$ follows
from \eqref{3.54}, and properties $(ii)$ and
$(iii)$ are implied by the orthogonality relations \eqref{3.56}, \eqref{3.57}, and the fact that
the matrix-valued Laurent polynomials $P_\pm$ and $R_\pm$ have invertible leading-order
coefficients (cf.\ Remark \ref{r2.4}).
\begin{corollary}
Let $k_0\in{\mathbb{Z}}$. Then the operators ${\mathbb{U}}_{\pm,k_0}$ are unitarily
equivalent to the operators of multiplication by $\zeta$ on
$\Ltm{\pm}$. In particular,
\begin{align}
& \sigma({\mathbb{U}}_{\pm,k_0}) = \text{\rm{supp}} \, (d\Omega_\pm(\cdot,k_0)).
\end{align}
\end{corollary}
\begin{proof}
Consider the linear maps $\dot
{\mathcal U}_\pm\colon\ltzm{[k_0,\pm\infty)\cap{\mathbb{Z}}}\to\Ltm{\pm}$ from the
space of compactly supported sequences
$\ltzm{[k_0,\pm\infty)\cap{\mathbb{Z}}}$ to the set of $\mathbb{C}^m$-valued Laurent
polynomials defined by
\begin{equation}
(\dot {\mathcal U}_\pm F)(z) = \sum_{k=k_0}^{\pm\infty} \widetilde
P_\pm(1/\overline{z},k,k_0)^* F(k), \quad F\in
\ltzm{[k_0,\pm\infty)\cap{\mathbb{Z}}}.
\end{equation}
Using \eqref{3.56} one shows that $\widehat F(\zeta) = (\dot {\mathcal U}_\pm
F)(\zeta)$, $F\in \ltzm{[k_0,\pm\infty)\cap{\mathbb{Z}}}$ has the
property
\begin{align}
\|\widehat F\|^2_{\Ltm{\pm}} &= \oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \widehat
F(\zeta)^*d\Omega_\pm(\zeta,k_0)\,\widehat F(\zeta)
\\ &=
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \sum_{k=k_0}^{\pm\infty} F(k)^*\widetilde P_\pm(\zeta,k,k_0)
\,d\Omega_\pm(\zeta,k_0)\!\! \sum_{k'=k_0}^{\pm\infty}\widetilde
P_\pm(\zeta,k',k_0)^*F(k') \notag
\\ &=
\sum_{k,k'=k_0}^{\pm\infty} F(k)^* \left(\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \widetilde
P_\pm(\zeta,k,k_0) \,d\Omega_\pm(\zeta,k_0)\, \widetilde
P_\pm(\zeta,k',k_0)^*\right)F(k') \notag
\\ &=
\sum_{k=k_0}^{\pm\infty} F(k)^*F(k) =
\|F\|^2_{\ltm{[k_0,\pm\infty)\cap{\mathbb{Z}}}}. \label{3.73}
\end{align}
Since $\ltzm{[k_0,\pm\infty)\cap{\mathbb{Z}}}$ is dense in
$\ltm{[k_0,\pm\infty)\cap{\mathbb{Z}}}$, $\dot {\mathcal U}_\pm$ extend to bounded
linear operators ${\mathcal U}_\pm\colon \ltm{[k_0,\pm\infty)\cap{\mathbb{Z}}} \to
\Ltm{\pm}$, and the identity
\begin{align} \label{3.74}
({\mathcal U}_\pm({\mathbb{U}}_{\pm,k_0}F))(\zeta) &= \sum_{k=k_0}^{\pm\infty} \widetilde
P_\pm(\zeta,k,k_0)^* ({\mathbb{U}}_{\pm,k_0}F)(k) =
\sum_{k=k_0}^{\pm\infty} ({\mathbb{U}}_{\pm,k_0}^*\widetilde
P_\pm(\zeta,\cdot,k_0))(k)^* F(k)
\\ &=
\sum_{k=k_0}^{\pm\infty} (\zeta^{-1}\widetilde P_\pm(\zeta,k,k_0))^* F(k)
= \zeta({\mathcal U}_\pm F)(\zeta), \quad F\in\ltm{[k_0,\pm\infty)\cap{\mathbb{Z}}},
\notag
\end{align}
holds. The ranges of the operators ${\mathcal U}_\pm$ are all of $\Ltm{\pm}$
since the sets of Laurent polynomials $\{\widetilde
P_\pm(\cdot,k,k_0)^*\}_{k\gtreqless k_0}$ are complete with respect
to $d\Omega_\pm(\cdot,k_0)$, and hence ${\mathcal U}_\pm$ are onto. Finally, one
computes the inverse operators ${\mathcal U}_\pm^{-1}$,
\begin{equation}
({\mathcal U}_\pm^{-1}\widehat F)(k) = \oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \widetilde
P_\pm(\zeta,k,k_0)\,d\Omega_{\pm}(\zeta,k_0)\,\widehat F(\zeta), \quad
\widehat F\in\Ltm{\pm},
\end{equation}
which together with \eqref{3.73} implies that ${\mathcal U}_\pm$ are unitary.
In addition, \eqref{3.74} implies that the half-lattice
unitary operators ${\mathbb{U}}_{\pm,k_0}$ on $\ltm{[k_0,\pm\infty)\cap{\mathbb{Z}}}$
are unitarily equivalent to the operators of multiplication by
$\zeta$ on $\Ltm{\pm}$,
\begin{align}
({\mathcal U}_\pm {\mathbb{U}}_{\pm,k_0} {\mathcal U}_\pm^{-1} \widehat F)(\zeta) = \zeta\widehat
F(\zeta), \quad \widehat F\in \Ltm{\pm}.
\end{align}
\end{proof}
\begin{corollary} \label{c3.8}
Let $k_0\in{\mathbb{Z}}$. \\ The matrix-valued Laurent polynomials
$\{P_+(\cdot,k,k_0)\}_{k\geq k_0}$ can be constructed by
Gram--Schmidt orthogonalizing
\begin{equation}
\begin{cases}
\zeta I_m,\, I_m,\,\zeta^2 I_m,\, \zeta^{-1}I_m,\, \zeta^3 I_m,\,
\zeta^{-2}I_m,\dots, & \text{$k_0$ odd,}
\\
I_m,\, \zeta I_m,\, \zeta^{-1}I_m,\, \zeta^2 I_m,\,
\zeta^{-2}I_m,\, \zeta^2 I_m, \dots, & \text{$k_0$ even}
\end{cases}
\end{equation}
in the context of matrix-valued Laurent polynomials orthogonal with respect
to $d\Omega_+(\cdot,k_0)$.
\\
The matrix-valued Laurent polynomials $\{R_+(\cdot,k,k_0)\}_{k\geq
k_0}$ can be constructed by Gram--Schmidt orthogonalizing
\begin{equation}
\begin{cases}
I_m,\, \zeta I_m,\, \zeta^{-1}I_m,\, \zeta^2 I_m,\,
\zeta^{-2}I_m,\, \zeta^3I_m,\dots, & \text{$k_0$ odd,}
\\
I_m,\, \zeta^{-1}I_m,\, \zeta I_m,\, \zeta^{-2}I_m,\,
\zeta^2 I_m,\, \zeta^{-3}I_m,\dots, & \text{$k_0$ even}
\end{cases}
\end{equation}
in the context of matrix-valued Laurent polynomials orthogonal with respect
to $d\Omega_+(\cdot,k_0)$.
\\
The matrix-valued Laurent polynomials $\{P_-(\cdot,k,k_0)\}_{k\leq
k_0}$ can be constructed by Gram--Schmidt orthogonalizing
\begin{equation}
\begin{cases}
I_m,\, -\zeta I_m,\, \zeta^{-1}I_m,\, -\zeta^2 I_m,\,
\zeta^{-2}I_m,\, -\zeta^3 I_m,\dots, & \text{$k_0$ odd,}
\\
-\zeta I_m,\, I_m,\, -\zeta^2 I_m,\, \zeta^{-1}I_m,\,
-\zeta^3I_m,\, \zeta^{-2}I_m, \dots, & \text{$k_0$ even}
\end{cases}
\end{equation}
in the context of matrix-valued Laurent polynomials orthogonal with respect
to $d\Omega_-(\cdot,k_0)$.
\\
The matrix-valued Laurent polynomials $\{R_-(\cdot,k,k_0)\}_{k\leq
k_0}$ can be constructed by Gram--Schmidt orthogonalizing
\begin{equation}
\begin{cases}
-I_m,\, \zeta^{-1}I_m,\, -\zeta I_m,\, \zeta^{-2}I_m,\,
-\zeta^2I_m,\, \zeta^{-3}I_m,\dots, & \text{$k_0$ odd,}
\\
I_m,\, -\zeta I_m,\, \zeta^{-1}I_m,\, -\zeta^2 I_m,\,
\zeta^{-2}I_m,\, -\zeta^3 I_m, \dots, & \text{$k_0$ even}
\end{cases}
\end{equation}
in the context of matrix-valued Laurent polynomials orthogonal with respect
to $d\Omega_-(\cdot,k_0)$.
Here the Gram--Schmidt orthogonalization procedure employs
left-multiplication by constant $($i.e., $\zeta$-independent\,$)$
$m\times m$ matrices as discussed in \cite[Sect.\ VII.2.8]{Be68}.
\end{corollary}
\begin{proof}
The result is a consequence of the definition of the Laurent polynomials
$P_\pm$ and $R_\pm$ and Lemma \ref{l3.6}.
\end{proof}
We note that the Gram--Schmidt orthogonalization
process implies that all matrix-valued Laurent polynomials constructed in Corollary \ref{c3.8} have
self-adjoint invertible leading-order coefficients (cf.\ Remark \ref{r2.4}).
The next result clarifies which measures arise as spectral
measures of half-lattice CMV operators and it yields the
reconstruction of the matrix-valued Verblunsky coefficients from the spectral measures
and the corresponding orthogonal Laurent polynomials.
\begin{theorem} \label{t3.9}
Let $k_0\in{\mathbb{Z}}$ and $d\Omega_\pm(\cdot,k_0)$ be nonnegative finite
measures on ${\partial\hspace*{.2mm}\mathbb{D}}$, supported on infinite sets, and normalized by
\begin{align}
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} d\Omega_\pm(\zeta,k_0) = I_m.
\end{align}
Moreover, assume that $d\Omega_\pm(\cdot,k_0)$ are nondegenerate in the
sense that expressions of the form
\begin{align}
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} P(\zeta) d\Omega_\pm(\zeta,k_0) P(\zeta)^*
\end{align}
are invertible for all ${\mathbb{C}^{m\times m}}$-valued Laurent polynomials $P(z)=z^{-n}A_{-n}+...+z^n A_n$ with either $A_{-n}=I_m$ or $A_{n}=I_m$. Then $d\Omega_\pm(\cdot,k_0)$ are necessarily
the spectral measures for some half-lattice CMV operators
${\mathbb{U}}_{\pm,k_0}$ with coefficients $\{\alpha_k\}_{k \geq k_0+1}$,
respectively, $\{\alpha_k\}_{k \leq k_0}$, defined by
\begin{equation}
\alpha_k = -
\begin{cases}
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \zeta
R_+(\zeta,k-1,k_0)d\Omega_+(\zeta,k_0)P_+(\zeta,k-1,k_0)^*, & k
\text{ odd,}
\\
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}
P_+(\zeta,k-1,k_0)d\Omega_+(\zeta,k_0)R_+(\zeta,k-1,k_0)^*, & k
\text{ even}
\end{cases} \label{3.83}
\end{equation}
for all $k \geq k_0+1$, and
\begin{equation}
\alpha_k = -
\begin{cases}
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \zeta
R_-(\zeta,k-1,k_0)d\Omega_-(\zeta,k_0)P_-(\zeta,k-1,k_0)^*, & k
\text{ odd,}
\\
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}
P_-(\zeta,k-1,k_0)d\Omega_-(\zeta,k_0)R_-(\zeta,k-1,k_0)^*, & k
\text{ even}
\end{cases} \label{3.84}
\end{equation}
for all $k \leq k_0$. Here the matrix-valued Laurent polynomials
$\{P_\pm(\cdot,k,k_0)\}_{k\geq k_0}$ and
$\{R_\pm(\cdot,k,k_0)\}_{k\geq k_0}$ denote the orthonormal
Laurent polynomials constructed in Corollary \ref{c3.8}.
\end{theorem}
\begin{proof}
First, using Corollary \ref{c3.8}, one constructs the orthonormal
polynomials $\{P_+(\cdot,k,k_0)\}_{k\geq k_0}$ and
$\{R_+(\cdot,k,k_0)\}_{k\geq k_0}$.
Next, we will establish the recursion relation \eqref{3.37}. Assume
$k$ is odd and consider the matrix-valued Laurent
polynomials $P$ and $R$,
\begin{align}
P(\zeta) &= \widetilde\rho_k P_+(\zeta,k,k_0) - \zeta
R_+(\zeta,k-1,k_0),
\\
R(\zeta) &= \rho_k R_+(\zeta,k,k_0) - \zeta^{-1}
P_+(\zeta,k-1,k_0),
\end{align}
where $\rho_k$, $\widetilde\rho_k \in {\mathbb{C}^{m\times m}}$ are self-adjoint invertible
matrices chosen such that the leading-order terms of the Laurent polynomials
$\widetilde\rho_k P_+(\zeta,k,k_0)$ and $\rho_k R_+(\zeta,k,k_0)$
cancel the leading-order terms of $\zeta R_+(\zeta,k-1,k_0)$ and
$\zeta^{-1} P_+(\zeta,k-1,k_0)$, respectively (cf.\ Remark \ref{r2.4}). Using Corollary
\ref{c3.8} one then checks that the Laurent polynomials $P$ and $R$ are
constant $m\times m$ matrix left-multiples of $P_+(\cdot,k-1,k_0)$
and $R_+(\cdot,k-1,k_0)$, respectively,
\begin{align}
\alpha_k P_+(\zeta,k-1,k_0) &= \widetilde\rho_k P_+(\zeta,k,k_0) -
\zeta R_+(\zeta,k-1,k_0), \label{3.87}
\\
\widetilde\alpha_k R_+(\zeta,k-1,k_0) &= \rho_k R_+(\zeta,k,k_0) -
\zeta^{-1} P_+(\zeta,k-1,k_0), \label{3.88}
\end{align}
with $\alpha_k$, $\widetilde\alpha_k \in {\mathbb{C}^{m\times m}}$ constant $m\times m$ matrices.
Moreover, using \eqref{3.87}, \eqref{3.88}, and Lemma \ref{l3.6} one
computes,
\begin{align}
I_m &= \oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \zeta R_+(\zeta,k-1,k_0)\,
d\Omega_\pm(\zeta,k_0)\, \zeta^{-1}R_+(\zeta,k-1,k_0)^* \notag
\\ &=
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \Big(\widetilde\rho_k P_+(\zeta,k,k_0)-\alpha_k
P_+(\zeta,k-1,k_0)\big)d\Omega_\pm(\zeta,k_0) \notag
\\ &\hspace*{2.45cm}
\times \big(\widetilde\rho_k P_+(\zeta,k,k_0)-\alpha_k
P_+(\zeta,k-1,k_0)\big)^* \notag
\\ &=
\widetilde\rho_k^2 + \alpha_k\alpha_k^*, \label{3.89}
\\
I_m &= \oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \zeta^{-1}P_+(\zeta,k-1,k_0)\,
d\Omega_\pm(\zeta,k_0)\, \zeta P_+(\zeta,k-1,k_0)^* \notag
\\ &=
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \Big(\rho_k R_+(\zeta,k,k_0)-\widetilde\alpha_k
R_+(\zeta,k-1,k_0)\big)d\Omega_\pm(\zeta,k_0) \notag
\\ &\hspace*{2.45cm}
\times \big(\rho_k R_+(\zeta,k,k_0)-\widetilde\alpha_k
R_+(\zeta,k-1,k_0)\big)^* \notag
\\ &=
\rho_k^2 + \widetilde\alpha_k\widetilde\alpha_k^*, \label{3.90}
\end{align}
and
\begin{align}
\alpha_k &= \oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \alpha_k P_+(\zeta,k-1,k_0)\,
d\Omega_\pm(\zeta,k_0)\,P_+(\zeta,k-1,k_0)^* \notag
\\ &=
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \Big(\widetilde\rho_k P_+(\zeta,k,k_0) -
\zeta R_+(\zeta,k-1,k_0)\big)d\Omega_\pm(\zeta,k_0)\,
P_+(\zeta,k-1,k_0)^* \notag
\\ &=
- \oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \zeta R_+(\zeta,k-1,k_0)\,d\Omega_\pm(\zeta,k_0)\,
P_+(\zeta,k-1,k_0)^*, \label{3.91}
\\
\widetilde\alpha_k &= \oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \widetilde\alpha_k R_+(\zeta,k-1,k_0)\,
d\Omega_\pm(\zeta,k_0)\,R_+(\zeta,k-1,k_0)^* \notag
\\ &=
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \Big(\rho_k R_+(\zeta,k,k_0) -
\zeta^{-1}P_+(\zeta,k-1,k_0)\big)d\Omega_\pm(\zeta,k_0)\,
R_+(\zeta,k-1,k_0)^* \notag
\\ &=
- \oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \zeta^{-1}P_+(\zeta,k-1,k_0)\,d\Omega_\pm(\zeta,k_0)\,
R_+(\zeta,k-1,k_0)^*. \label{3.92}
\end{align}
Thus, \eqref{3.89}--\eqref{3.92} imply that $\widetilde\alpha_k =
\alpha_k^*$, $\rho_k = \sqrt{I_m-\alpha_k^*\alpha_k}$, and $\widetilde\rho_k =
\sqrt{I_m-\alpha_k\alpha_k^*}$, and hence \eqref{3.87} and \eqref{3.88}
yield the recursion relation \eqref{3.37}. A similar argument also
proves the recursion relation \eqref{3.37} for the case $k$ even.
Finally, using Lemma \ref{l3.3} one concludes that the Laurent
polynomials $\{P_+(\cdot,k,k_0)\}_{k\geq k_0}$ form a generalized
eigenvector of the operator ${\mathbb{U}}_{+,k_0}$ associated with the
coefficients $\alpha_k,\rho_k,\widetilde\rho_k$ introduced above. Thus, the
measure $d\Omega_+(\cdot,k_0)$ is the spectral measure of the operator
${\mathbb{U}}_{+,k_0}$.
Similarly one proves the result for $d\Omega_-(\cdot,k_0)$ and
\eqref{3.84} for $k\leq k_0$.
\end{proof}
\begin{lemma} \label{l3.10}
Let $z\in\mathbb{C}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\})$ and $k_0\in{\mathbb{Z}}$. Then the
following identity holds,
\begin{align}
\begin{split} \label{3.93}
& \widetilde Q_\pm(z,k,k_0) = \pm \oint_{{\partial\hspace*{.2mm}\mathbb{D}}}
\frac{\zeta+z}{\zeta-z} \big(\widetilde P_\pm(\zeta,k,k_0) - \widetilde
P_\pm(z,k,k_0)\big)d\Omega_{\pm}(\zeta,k_0), \quad k \gtrless k_0,
\\
&S_\pm(z,k,k_0) = \pm \oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \frac{\zeta+z}{\zeta-z}
\big(R_\pm(\zeta,k,k_0)-R_\pm(z,k,k_0)\big) d\Omega_{\pm}(\zeta,k_0),
\quad k \gtrless k_0.
\end{split}
\end{align}
\end{lemma}
\begin{proof}
To simplify our further notation we agree to write both equalities in
\eqref{3.93} as a single one,
\begin{align} \label{3.95}
&\binom{\widetilde Q_\pm(z,k,k_0)}{S_\pm(z,k,k_0)} = \pm\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}
\frac{\zeta+z}{\zeta-z} \bigg( \binom{\widetilde
P_\pm(\zeta,k,k_0)}{R_\pm(\zeta,k,k_0)} - \binom{\widetilde
P_\pm(z,k,k_0)}{R_\pm(z,k,k_0)} \bigg) d\Omega_{\pm}(\zeta,k_0), \quad
k\gtrless k_0,
\end{align}
where the integration on the right-hand side is understood
componentwise, that is, an expression of the type $\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}
\left(\begin{smallmatrix} G_1(\zeta) \\ G_2(\zeta)\end{smallmatrix}\right)
d\Omega_\pm(\zeta,k_0)$
with $G_1(z)$ and $G_2(z)$ some ${\mathbb{C}^{m\times m}}$-valued Laurent polynomials is defined by
\begin{align}
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \binom{G_1(\zeta)}{G_2(\zeta)}d\Omega_\pm(\zeta,k_0) =
\begin{pmatrix} \oint_{{\partial\hspace*{.2mm}\mathbb{D}}}G_1(\zeta) \, d\Omega_\pm(\zeta,k_0) \\
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}G_2(\zeta) \, d\Omega_\pm(\zeta,k_0) \end{pmatrix}.
\end{align}
First, we prove \eqref{3.95} for the right half-lattice Laurent polynomials
and for $k_0$ even. In this case \eqref{3.49a} and
\eqref{3.50a} imply that \eqref{3.95} is equivalent to
\begin{align}
&\binom{Q_+(z,k,k_0)}{S_+(z,k,k_0)} = \oint_{{\partial\hspace*{.2mm}\mathbb{D}}}
\frac{\zeta+z}{\zeta-z} \bigg(
\binom{P_+(\zeta,k,k_0)}{R_+(\zeta,k,k_0)} -
\binom{P_+(z,k,k_0)}{R_+(z,k,k_0)} \bigg) d\Omega_{+}(\zeta,k_0), \quad k
> k_0, \text{ $k_0$ even}. \label{3.97}
\end{align}
Let $k_0\in{\mathbb{Z}}$ be even. It suffices to show that the right-hand side
of \eqref{3.97}, temporarily denoted by the symbol $RHS(z,k,k_0)$,
satisfies
\begin{align}
& {\mathbb{T}}(z,k+1)^{-1} RHS(z,k+1,k_0) = RHS(z,k,k_0),
\quad k > k_0, \label{3.98} \\
& {\mathbb{T}}(z,k_0+1)^{-1}RHS(z,k_0+1,k_0) =
\binom{Q_+(z,k_0,k_0)}{S_+(z,k_0,k_0)} = \binom{-I_m}{I_m}. \label{3.99}
\end{align}
One verifies these statements using the equality,
\begin{align} \label{3.100}
& {\mathbb{T}}(z,k+1)^{-1} RHS(z,k+1,k_0) = RHS(z,k,k_0)
\\ \notag
& \quad + \oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \frac{\zeta+z}{\zeta-z} \left(
{\mathbb{T}}(z,k+1)^{-1} - {\mathbb{T}}(\zeta,k+1)^{-1}\right)
\binom{P_+(\zeta,k+1,k_0)}{R_+(\zeta,k+1,k_0)}d\Omega_{+}(\zeta,k_0),
\quad k\in{\mathbb{Z}}.
\end{align}
For $k>k_0$, the last term on the right-hand side of \eqref{3.100}
vanishes since for $k$ odd, ${\mathbb{T}}(z,k+1)$ does not depend on $z$, and
for $k$ even, it follows from Corollary \ref{c3.8} that
$P_+(\cdot,k+1,k_0)$ and $R_+(\cdot,k+1,k_0)$ are orthogonal in
$\Ltm{+}$ to ${\text{\rm span}}\{I_m,\zeta I_m\}$ and
${\text{\rm span}}\{I_m,\zeta^{-1} I_m\}$, respectively. Hence one computes
\begin{align}
&\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \frac{\zeta+z}{\zeta-z} \Big({\mathbb{T}}(z,k+1)^{-1}-
{\mathbb{T}}(\zeta,k+1)^{-1}\Big)
\binom{P_+(\zeta,k+1,k_0)}{R_+(\zeta,k+1,k_0)}d\Omega_{+}(\zeta,k_0)
\notag
\\
& \quad = \oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \frac{\zeta+z}{\zeta-z}
\begin{pmatrix}
0 & (z-\zeta)\rho_{k+1}^{-1} \\
(z^{-1} - \zeta^{-1})\widetilde\rho_{k+1}^{-1} & 0
\end{pmatrix}
\binom{P_+(\zeta,k+1,k_0)}{R_+(\zeta,k+1,k_0)}d\Omega_{+}(\zeta,k_0)
\notag
\\
& \quad = \oint_{{\partial\hspace*{.2mm}\mathbb{D}}}
\begin{pmatrix}
0 & -(\zeta+z)\rho_{k+1}^{-1} \\
(\zeta^{-1} + z^{-1})\widetilde\rho_{k+1}^{-1} & 0
\end{pmatrix}
\binom{P_+(\zeta,k+1,k_0)}{R_+(\zeta,k+1,k_0)}d\Omega_{+}(\zeta,k_0) \notag
\\ & \quad = \oint_{{\partial\hspace*{.2mm}\mathbb{D}}}
\binom{-\rho_{k+1}^{-1}(\zeta+z)R_+(\zeta,k,k_0)}
{\widetilde\rho_{k+1}^{-1}(\zeta^{-1}+z^{-1})P_+(\zeta,k,k_0)}
d\Omega_{+}(\zeta,k_0) = \binom{0}{0}.
\end{align}
Thus, \eqref{3.98} is implied by \eqref{3.100}.
For $k=k_0$ even, one obtains that $RHS(z,k_0,k_0)=0$ since by \eqref{3.49} one has the normalization
$P_+(z,k_0,k_0)=R_+(z,k_0,k_0)=I_m$. Then using the fact that by
Corollary \ref{c3.8}, $P_+(\cdot,k_0+1,k_0)$ and
$R_+(\cdot,k_0+1,k_0)$ are orthogonal to constant $m\times m$
matrices in $\Ltm{+}$ and that by \eqref{Table},
\begin{align}
\begin{split}
P_+(\zeta,k_0+1,k_0) &=
\widetilde\rho_{k_0+1}^{-1}(\zeta I_m+\alpha_{k_0+1}),\\
R_+(\zeta,k_0+1,k_0) &=
\rho_{k_0+1}^{-1}(\zeta^{-1}I_m+\alpha_{k_0+1}^*),
\end{split}
\end{align}
one computes,
\begin{align}
&\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \zeta^{-1}P_+(\zeta,k_0+1,k_0)d\Omega_+(\zeta,k_0) =
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} P_+(\zeta,k_0+1,k_0)d\Omega_+(\zeta,k_0) (\zeta I_m)^*
\notag
\\ &\quad =
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} P_+(\zeta,k_0+1,k_0)d\Omega_+(\zeta,k_0)
P_+(\zeta,k_0+1,k_0)^*\widetilde\rho_{k_0+1} = \widetilde\rho_{k_0+1},
\label{3.104}
\\
&\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \zeta R_+(\zeta,k_0+1,k_0)d\Omega_+(\zeta,k_0) =
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} R_+(\zeta,k_0+1,k_0)d\Omega_+(\zeta,k_0) (\zeta^{-1}I_m)^*
\notag
\\ &\quad =
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} R_+(\zeta,k_0+1,k_0)d\Omega_+(\zeta,k_0)
R_+(\zeta,k_0+1,k_0)^*\rho_{k_0+1} = \rho_{k_0+1}, \label{3.105}
\end{align}
and hence,
\begin{align} \label{3.106}
&\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \frac{\zeta+z}{\zeta-z}
\Big({\mathbb{T}}(z,k_0+1)^{-1}-{\mathbb{T}}(\zeta,k_0+1)^{-1}\Big)
\binom{P_+(\zeta,k_0+1,k_0)}{R_+(\zeta,k_0+1,k_0)}d\Omega_{+}(\zeta,k_0)
\notag
\\
& \quad = \oint_{{\partial\hspace*{.2mm}\mathbb{D}}}
\binom{-\rho_{k+1}^{-1}(\zeta+z)R_+(\zeta,k_0+1,k_0)}
{\widetilde\rho_{k+1}^{-1}(\zeta^{-1}+z^{-1})P_+(\zeta,k_0+1,k_0)}
d\Omega_{+}(\zeta,k_0)
\\ & \quad =
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \binom{-\rho_{k+1}^{-1}\zeta R_+(\zeta,k_0+1,k_0)}
{\widetilde\rho_{k+1}^{-1}\zeta^{-1}P_+(\zeta,k_0+1,k_0)}
d\Omega_{+}(\zeta,k_0) =
\binom{-\rho_{k+1}^{-1}\rho_{k+1}}{\widetilde\rho_{k+1}^{-1}\widetilde\rho_{k+1}}
= \binom{-I_m}{I_m}. \notag
\end{align}
Thus, \eqref{3.99} is a consequence of \eqref{3.100}, \eqref{3.106}, and the
fact that $RHS(z,k_0,k_0)=0$.
Next, we prove \eqref{3.95} for the right half-lattice Laurent polynomials
and $k_0$ odd,
\begin{align}
&\binom{S_+(z,k,k_0)}{\widetilde Q_+(z,k,k_0)} = \oint_{{\partial\hspace*{.2mm}\mathbb{D}}}
\frac{\zeta+z}{\zeta-z} \bigg(\binom{R_+(\zeta,k,k_0)}{\widetilde
P_+(\zeta,k,k_0)} - \binom{R_+(z,k,k_0)}{\widetilde P_+(z,k,k_0)} \bigg)
d\Omega_{+}(\zeta,k_0), \quad k > k_0, \text{ $k_0$ odd}. \label{3.107}
\end{align}
Let $k_0\in{\mathbb{Z}}$ be odd. We note that for $U(z,k),V(z,k)\in{\mathbb{C}^{m\times m}}$,
$k\in{\mathbb{Z}}$, $z\in{\C\backslash\{0\}}$,
\begin{align}
\binom{U(z,k)}{V(z,k)} &= {\mathbb{T}}(z,k) \binom{U(z,k-1)}{V(z,k-1)}
\\
\intertext{is equivalent to } \binom{V(z,k)}{\widetilde U(z,k)} &= \widetilde
{\mathbb{T}}(z,k) \binom{V(z,k-1)}{\widetilde U(z,k-1)},
\end{align}
where
\begin{equation}
\widetilde U(z,k) = z^{-1} U(z,k) \,\text{ and }\, \widetilde {\mathbb{T}}(z,k) =
\begin{pmatrix}0 & I_m \\ z^{-1}I_m & 0\end{pmatrix} {\mathbb{T}}(z,k)
\begin{pmatrix}0 & zI_m \\ I_m & 0\end{pmatrix}.
\end{equation}
Thus, it suffices to show that the right-hand side of \eqref{3.107},
temporarily denoted by $\widetilde{RHS}(z,k,k_0)$, satisfies
\begin{align}
& \widetilde {\mathbb{T}}(z,k+1)^{-1} \widetilde{RHS} (z,k+1,k_0) = \widetilde{RHS}(z,k,k_0),
\quad k > k_0,
\\
& \widetilde {\mathbb{T}}(z,k_0+1)^{-1} \widetilde{RHS}(z,k_0+1,k_0) =
\binom{S_+(z,k_0,k_0)}{\widetilde Q_+(z,k_0,k_0)} = \binom{-I_m}{I_m}.
\end{align}
At this point one can follow the first part of the proof replacing
${\mathbb{T}}$ by $\widetilde {\mathbb{T}}$, $\Big(\begin{smallmatrix} P_+ \\[1mm]
R_+\end{smallmatrix}\Big)$ by $\Big(\begin{smallmatrix} R_+ \\[.5mm]
\widetilde P_+\end{smallmatrix}\Big)$, $\Big(\begin{smallmatrix} Q_+
\\[1mm] S_+\end{smallmatrix}\Big)$ by $\Big(\begin{smallmatrix}
S_+ \\[.5mm] \widetilde Q_+\end{smallmatrix}\Big)$, etc.
The result for the remaining Laurent polynomials $\widetilde P_-(z,k,k_0)$,
$R_-(z,k,k_0)$, $\widetilde Q_-(z,k,k_0)$, and $S_-(z,k,k_0)$ is proved
similarly.
\end{proof}
\begin{lemma} \label{l3.11}
Let $k_0\in{\mathbb{Z}}$ and let $m_\pm(\cdot,k_0)$ denote the
${\mathbb{C}^{m\times m}}$-valued Caratheodory and anti-Caratheodory functions
\begin{align}
m_\pm(z,k_0) &= \pm \Delta_{k_0}^*({\mathbb{U}}_{\pm,k_0}+zI)({\mathbb{U}}_{\pm,k_0}-zI)^{-1}
\Delta_{k_0} \label{3.110}
\\
& =\pm
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}d\Omega_{\pm}(\zeta,k_0)\,\frac{\zeta+z}{\zeta-z},
\quad z\in{\mathbb{C}}\backslash{\partial\hspace*{.2mm}\mathbb{D}}, \label{3.111}
\end{align}
with
\begin{equation}
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}d\Omega_{\pm}(\zeta,k_0)= I_m. \label{3.111a}
\end{equation}
Then the following relations hold,
\begin{align}
Q_\pm(z,\cdot,k_0) + P_\pm(z,\cdot,k_0)m_\pm(z,k_0) \in
\ltmm{[k_0,\pm\infty)\cap{\mathbb{Z}}}, \quad z\in{\mathbb{C}}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\}),
\label{3.112}
\\
S_\pm(z,\cdot,k_0) + R_\pm(z,\cdot,k_0)m_\pm(z,k_0) \in
\ltmm{[k_0,\pm\infty)\cap{\mathbb{Z}}}, \quad z\in{\mathbb{C}}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\}).
\label{3.113}
\end{align}
\end{lemma}
\begin{proof}
Equality \eqref{3.111} is implied by \eqref{3.54} and \eqref{3.55}. \\
Next, let ${\mathbb{B}}_{\pm,k_0}(z)$ denote operators defined on
$\ltm{[k_0,\pm\infty)\cap{\mathbb{Z}}}$ by
\begin{align}
{\mathbb{B}}_{\pm,k_0}(z) = ({\mathbb{U}}_{\pm,k_0}+zI)({\mathbb{U}}_{\pm,k_0}-zI)^{-1}, \quad
z\in{\mathbb{C}}\backslash{\partial\hspace*{.2mm}\mathbb{D}}.
\end{align}
Since ${\mathbb{U}}_{\pm,k_0}$ are unitary, the operators ${\mathbb{B}}_{\pm,k_0}(z)$
are bounded for all $z\in{\mathbb{C}}\backslash{\partial\hspace*{.2mm}\mathbb{D}}$, and hence one has
\begin{align}
{\mathbb{B}}_{\pm,k_0}(z)\Delta_{k_0}=
\big\{\Delta_{k}^*{\mathbb{B}}_{\pm,k_0}(z)\Delta_{k_0}\big\}_{k\in[k_0,\pm\infty)\cap{\mathbb{Z}}}
\in\ltmm{[k_0,\pm\infty)\cap{\mathbb{Z}}}. \label{3.115}
\end{align}
Using the spectral representation for the operators
${\mathbb{B}}_{\pm,k_0}(z)$ and equalities \eqref{3.64}, \eqref{3.93}, and \eqref{3.111}, one obtains
\begin{align}
\Delta_{k}^*{\mathbb{B}}_{\pm,k_0}(z)\Delta_{k_0} &=
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}\frac{\zeta+z}{\zeta-z} \widetilde
P_\pm(\zeta,k,k_0) \, d\Omega_{\pm}(\zeta,k_0) \notag
\\
&= \pm\big[\widetilde Q_\pm(z,k,k_0)+ \widetilde P_\pm(z,k,k_0)
m_\pm(z,k_0)\big], \quad k \gtrless k_0. \label{3.116}
\end{align}
Thus, \eqref{3.112} is a consequence of \eqref{3.49a}, \eqref{3.50a},
\eqref{3.115}, and \eqref{3.116}.
Moreover, since $\Big(\begin{smallmatrix}P_\pm(z,k,k_0)\\[1mm]
R_\pm(z,k,k_0)\end{smallmatrix}\Big)_{k\in{\mathbb{Z}}}$ and
$\Big(\begin{smallmatrix} Q_\pm(z,k,k_0)\\[1mm] S_\pm(z,k,k_0)
\end{smallmatrix}\Big)_{k\in{\mathbb{Z}}}$,
$z\in{\mathbb{C}}\backslash\{0\}$, satisfy \eqref{3.21}, Lemma \ref{l3.2} implies that
\begin{align}
({\mathbb{W}} (Q_\pm (z,\cdot,k_0)+P_\pm (z,\cdot,k_0)m_\pm(z,k_0)))(k) =
z[S_\pm (z,k,k_0)+R_\pm (z,k,k_0)m_\pm(z,k_0)], \quad k\in{\mathbb{Z}}, \label{3.117}
\end{align}
and hence \eqref{3.113} follows from \eqref{3.112} and \eqref{3.117}.
\end{proof}
\begin{lemma} \label{l3.12}
Let $k_0\in{\mathbb{Z}}$. Then the relations in \eqref{3.112} $($equivalently, those in
\eqref{3.113}$)$ uniquely determine the ${\mathbb{C}^{m\times m}}$-valued functions
$m_\pm(\cdot,k_0)$ on ${\mathbb{C}}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\})$.
\end{lemma}
\begin{proof}
We will prove the lemma by contradiction. Assume that there are two
${\mathbb{C}^{m\times m}}$-valued functions $m_+(z,k_0)$ and $\widetilde m_+(z,k_0)$ satisfying
\eqref{3.112} such that $m_+(z_0,k_0) \neq \widetilde m_+(z_0,k_0)$ for
some $z_0\in\mathbb{C}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\})$. Then there is a vector
$x\in\mathbb{C}^m$ such that $(m_+(z_0,k_0) - \widetilde m_+(z_0,k_0))x \neq 0$ and
by \eqref{3.112},
\begin{align}
p_+(z_0,\cdot,k_0)=P_+(z_0,\cdot,k_0)[m_+(z_0,k_0)-\widetilde
m_+(z_0,k_0)]x \in \ltm{[k_0,\pm\infty)\cap{\mathbb{Z}}}, \quad
z\in{\mathbb{C}}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\}).
\end{align}
Since $P_+(z_0,k_0,k_0)\neq 0$, the sequence of vectors
$\{p_+(z,k,k_0)\}_{k\geq k_0}$ is not identically zero, and hence, by
Lemma \ref{l3.3}, $p_+(z_0,\cdot,k_0)$ is an eigenvector of the
operator ${\mathbb{U}}_{+,k_0}$ corresponding to the eigenvalue
$z_0\in\mathbb{C}\backslash{\partial\hspace*{.2mm}\mathbb{D}}$. This contradicts unitarity of ${\mathbb{U}}_{+,k_0}$.
Similarly, one proves the result for $m_-(z,k_0)$. Moreover, one
easily supplies a proof that utilizes \eqref{3.113} instead of
\eqref{3.112}.
\end{proof}
\begin{corollary} \label{c3.13}
There are solutions $\Big(\begin{smallmatrix}\psi_\pm(z,k)\\[1mm]
\chi_\pm(z,k)\end{smallmatrix}\Big)$, $k\in{\mathbb{Z}}$, of \eqref{3.21},
unique up to right-multiplication by constant $m\times m$ matrices, so that for some $($and hence for
all\,$)$ $k_1\in{\mathbb{Z}}$,
\begin{align}
\psi_\pm(z,\cdot),\,\chi_\pm(z,\cdot) \in
\ell^2 ([k_1,\pm\infty)\cap{\mathbb{Z}})^{m\times m}, \quad
z\in\mathbb{C}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\}).
\end{align}
\end{corollary}
\begin{proof}
Since any solution of \eqref{3.21} can be expressed as a linear
combination of the Laurent polynomials
$\Big(\begin{smallmatrix}P_\pm(z,\cdot,k_0)\\[1mm]
R_\pm(z,\cdot,k_0)\end{smallmatrix}\Big)$ and
$\Big(\begin{smallmatrix}Q_\pm(z,\cdot,k_0)\\[1mm]
S_\pm(z,\cdot,k_0)\end{smallmatrix}\Big)$, existence and uniqueness
of the solutions
$\Big(\begin{smallmatrix}\psi_\pm(z,\cdot)\\[1mm]
\chi_\pm(z,\cdot)\end{smallmatrix}\Big)$ is implied by Lemmas
\ref{l3.11} and \ref{l3.12}, respectively.
\end{proof}
For the next result we recall the definition of $a_k$ and $b_k$ in \eqref{3.10}
and \eqref{3.11}.
\begin{lemma} \label{l3.14}
Let $z\in{\mathbb{C}}\backslash\{0\}$ and $k_0\in{\mathbb{Z}}$. Then the following
relations hold for all $k\in{\mathbb{Z}}$,
\begin{align}
\binom{P_-(z,k,k_0-1)}{R_-(z,k,k_0-1)} &=
\binom{P_+(z,k,k_0)}{R_+(z,k,k_0)}
\frac{\widetilde\rho_{k_0}^{-1}b_{k_0}-\rho_{k_0}^{-1}b_{k_0}^*}{2} +
\binom{Q_+(z,k,k_0)}{S_+(z,k,k_0)}
\frac{\widetilde \rho_{k_0}^{-1}b_{k_0}+\rho_{k_0}^{-1}b_{k_0}^*}{2}, \label{3.120}
\\
\binom{Q_-(z,k,k_0-1)}{S_-(z,k,k_0-1)} &=
\binom{P_+(z,k,k_0)}{R_+(z,k,k_0)}
\frac{\widetilde\rho_{k_0}^{-1}a_{k_0}+\rho_{k_0}^{-1}a_{k_0}^*}{2} +
\binom{Q_+(z,k,k_0)}{S_+(z,k,k_0)}
\frac{\widetilde \rho_{k_0}^{-1}a_{k_0}-\rho_{k_0}^{-1}a_{k_0}^*}{2}, \label{3.121}
\end{align}
and
\begin{align}
\binom{P_-(z,k,k_0)}{R_-(z,k,k_0)} &=
\binom{P_+(z,k,k_0)}{R_+(z,k,k_0)} c(z,k_0) +
\binom{Q_+(z,k,k_0)}{S_+(z,k,k_0)} d(z,k_0), \label{3.122}
\\
\binom{Q_-(z,k,k_0)}{S_-(z,k,k_0)} &=
\binom{P_+(z,k,k_0)}{R_+(z,k,k_0)} d(z,k_0) +
\binom{Q_+(z,k,k_0)}{S_+(z,k,k_0)} c(z,k_0), \label{3.123}
\end{align}
where
\begin{align} \label{3.124}
c(z,k_0) =
\begin{cases}
\frac{1-z}{2z}, & k_0 \text{ odd},\\
\frac{1-z}{2}, & k_0 \text{ even}
\end{cases}
\,\text{ and }\, d(z,k_0) =
\begin{cases}
\frac{1+z}{2z}, & k_0 \text{ odd},\\
\frac{1+z}{2}, & k_0 \text{ even}.
\end{cases}
\end{align}
\end{lemma}
\begin{proof}
Since the left and right-hand sides of \eqref{3.120}--\eqref{3.123}
satisfy the same recursion relation \eqref{3.21}, it suffices to
check \eqref{3.120}--\eqref{3.123} at only one point, say, at the
point $k=k_0$. The latter is easily seen to be a consequence of \eqref{Table}.
\end{proof}
\begin{theorem} \label{t3.15}
Let $k_0\in{\mathbb{Z}}$. Then there exist unique ${\mathbb{C}^{m\times m}}$-valued functions
$M_\pm(\cdot,k_0)$ such that for all
$z\in{\mathbb{C}}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\})$
\begin{align}
&U_\pm(z,\cdot,k_0) = Q_+(z,\cdot,k_0) + P_+(z,\cdot,k_0)M_\pm(z,k_0)
\in \ltmm{[k_0,\pm\infty)\cap{\mathbb{Z}}}, \label{3.125}
\\
&V_\pm(z,\cdot,k_0) = S_+(z,\cdot,k_0) + R_+(z,\cdot,k_0)M_\pm(z,k_0)
\in \ltmm{[k_0,\pm\infty)\cap{\mathbb{Z}}}. \label{3.126}
\end{align}
\end{theorem}
\begin{proof}
The assertions \eqref{3.125} and \eqref{3.126} follow from
Lemma \ref{l3.11}, Corollary \ref{c3.13}, and Lemma \ref{l3.14}.
\end{proof}
We will call $U_\pm(z,\cdot,k_0)$ the {\it Weyl--Titchmarsh solutions} of
${\mathbb{U}}$. By Corollary \ref{c3.13}, $U_\pm(z,\cdot,k_0)$ and
$V_\pm(z,\cdot,k_0)$ are unique up to right-multiplication by constant $m\times m$ matrices. Similarly,
we will call $m_\pm(z,k_0)$ as well as $M_\pm(z,k_0)$ the {\it
half-lattice Weyl--Titchmarsh $m$-functions} associated with
${\mathbb{U}}_{\pm,k_0}$. (See also \cite{Si04a} for a comparison of various
alternative notions of Weyl--Titchmarsh $m$-functions for
${\mathbb{U}}_{+,k_0}$ with scalar-valued Verblunsky coefficients.)
Lemma \ref{l3.11}, Corollary \ref{c3.13}, and Lemma
\ref{l3.14} imply that for all $z\in{\mathbb{C}}\backslash{\partial\hspace*{.2mm}\mathbb{D}}$,
\begin{align}
M_+(z,k_0) &= m_+(z,k_0), \label{3.127}
\\
M_+(0,k_0) &=I_m, \label{3.128}
\\
M_-(z,k_0) &=[(1+z)I_m + (1-z)m_-(z,k_0)][(1-z)I_m +
(1+z)m_-(z,k_0)]^{-1}, \label{3.129}
\\
M_-(z,k_0) &=
\big[(\widetilde\rho_{k_0}^{-1}a_{k_0}+\rho_{k_0}^{-1}a_{k_0}^*) +
(\widetilde\rho_{k_0}^{-1}b_{k_0}-\rho_{k_0}^{-1}b_{k_0}^*)m_-(z,k_0-1)\big]
\notag
\\
&\quad\times
\big[(\widetilde\rho_{k_0}^{-1}a_{k_0}-\rho_{k_0}^{-1}a_{k_0}^*) +
(\widetilde\rho_{k_0}^{-1}b_{k_0}+\rho_{k_0}^{-1}b_{k_0}^*)m_-(z,k_0-1)\big]^{-1},
\label{3.130}
\\
M_-(0,k_0) &=(\alpha_{k_0}+I_m)(\alpha_{k_0}-I_m)^{-1}. \label{3.131}
\end{align}
In addition, it follows from \eqref{3.111} and Theorem \ref{tA.2}
that $m_\pm(z,k_0)$ are ${\mathbb{C}^{m\times m}}$-valued Caratheodory and
anti-Caratheodory functions, respectively. From \eqref{3.127} one
concludes that $M_+(z,k_0)$ is also a Caratheodory function. Using
\eqref{3.129} one verifies that $M_-(z,k_0)$ is analytic in $\mathbb{D}$
since the anti-Caratheodory function $m_-(\cdot,k_0)$ satisfies
$\Re(m_-(z,k_0))=(m_-(z,k_0)+m_-(z,k_0)^*)/2 < 0$ for $z\in\mathbb{D}$.
Moreover, utilizing \eqref{3.12}, \eqref{3.13}, and \eqref{3.130},
one computes,
\begin{align}
\Re(M_-(z,k_0)) &= [M_-(z,k_0)+M_-(z,k_0)^*]/2 \notag
\\
&= \big[(a_{k_0}^*\widetilde\rho_{k_0}^{-1}-a_{k_0}\rho_{k_0}^{-1}) +
m_-(z,k_0-1)^*(b_{k_0}^*\widetilde\rho_{k_0}^{-1}+b_{k_0}\rho_{k_0}^{-1})\big]^{-1}
\Re(m_-(z,k_0-1)) \notag
\\
&\quad\times
\big[(\widetilde\rho_{k_0}^{-1}a_{k_0}-\rho_{k_0}^{-1}a_{k_0}^*) +
(\widetilde\rho_{k_0}^{-1}b_{k_0}+\rho_{k_0}^{-1}b_{k_0}^*)m_-(z,k_0-1)\big]^{-1},
\end{align}
and hence, $M_-(\cdot,k_0)$ is an anti-Caratheodory matrix.
Next, we introduce the ${\mathbb{C}^{m\times m}}$-valued functions $\Phi_\pm(\cdot,k)$,
$k\in{\mathbb{Z}}$, by
\begin{align}
\Phi_\pm(z,k) = [M_\pm(z,k)-I_m][M_\pm(z,k)+I_m]^{-1}, \quad
z\in\mathbb{C}\backslash{\partial\hspace*{.2mm}\mathbb{D}}. \label{3.133}
\end{align}
Then \eqref{3.128} and \eqref{3.131} imply that
\begin{equation} \label{3.133a}
\Phi_+(0,k_0) = 0 \,\text{ and }\, \Phi_-(0,k_0)^{-1}=\alpha_{k_0}.
\end{equation}
Moreover, one verifies that
\begin{align}
M_\pm(z,k) &= [I_m-\Phi_\pm(z,k)]^{-1}[I_m+\Phi_\pm(z,k)], \quad
z\in\mathbb{C}\backslash{\partial\hspace*{.2mm}\mathbb{D}}, \label{3.134}
\\
m_-(z,k) &= [zI_m+\Phi_-(z,k)]^{-1}[zI_m-\Phi_-(z,k)], \quad
z\in\mathbb{C}\backslash{\partial\hspace*{.2mm}\mathbb{D}} \label{3.135}
\end{align}
(cf.\ Remark \ref{r3.18}). In addition, we extend these functions to the unit circle ${\partial\hspace*{.2mm}\mathbb{D}}$ by
taking the radial limits which exist and are finite for Lebesgue
almost every $\zeta\in{\partial\hspace*{.2mm}\mathbb{D}}$,
\begin{align}
M_\pm(\zeta,k) &= \lim_{r \uparrow 1} M_\pm(r\zeta,k),
\\
\Phi_\pm(\zeta,k) &= \lim_{r \uparrow 1} \Phi_\pm(r\zeta,k), \quad
k\in{\mathbb{Z}}.
\end{align}
\begin{lemma} \label{l3.16}
Let $z\in\mathbb{C}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\})$, $k_0, k\in{\mathbb{Z}}$. Then the
functions $\Phi_\pm(\cdot,k)$ satisfy
\begin{equation}
\Phi_\pm(z,k) =
\begin{cases}
zV_\pm(z,k,k_0)U_\pm(z,k,k_0)^{-1}, &\text{$k$ odd,}
\\
U_\pm(z,k,k_0)V_\pm(z,k,k_0)^{-1}, & \text{$k$ even,}
\end{cases}
\end{equation}
where $U_\pm(\cdot,k,k_0)$ and $V_\pm(\cdot,k,k_0)$ are the
${\mathbb{C}^{m\times m}}$-valued functions defined in \eqref{3.125} and \eqref{3.126},
respectively.
\end{lemma}
\begin{proof}
Using Corollary \ref{c3.13} it suffices to assume $k=k_0$. Then the
statement is immediately implied by \eqref{3.49}, \eqref{3.125},
\eqref{3.126}, and \eqref{3.133}.
\end{proof}
\begin{lemma} \label{l3.17}
Let $k\in{\mathbb{Z}}$. Then the ${\mathbb{C}^{m\times m}}$-valued functions
$\Phi_+(\cdot,k)|_{\mathbb{D}}$ $($resp., $\Phi_-(\cdot,k)|_{\mathbb{D}}$$)$ are
Schur $($resp., anti-Schur\,$)$ matrices. Moreover, $\Phi_\pm$
satisfy the Riccati-type equation
\begin{equation}
\Phi_\pm(z,k)\widetilde\rho_k^{-1}\alpha_k\Phi_\pm(z,k-1) +
z\Phi_\pm(z,k)\widetilde\rho_k^{-1} - \rho_k^{-1}\Phi_\pm(z,k-1)
=z\rho_k^{-1}\alpha_k^*, \quad z\in{\mathbb{C}}\backslash{\partial\hspace*{.2mm}\mathbb{D}},\; k\in{\mathbb{Z}}.
\label{3.138}
\end{equation}
\end{lemma}
\begin{proof}
Lemma \ref{l3.16} and \eqref{3.133} imply that the
functions $\Phi_+(\cdot,k)|_{\mathbb{D}}$ $($resp.,
$\Phi_-(\cdot,k)|_{\mathbb{D}}$$)$ are Schur $($resp., anti-Schur\,$)$
matrices.
Let $k$ be odd. Then applying Lemma \ref{l3.16} and the recursion
relation \eqref{3.21} one obtains
\begin{align}
&\Phi_\pm(z,k) = zV_\pm(z,k,k_0)U_\pm(z,k,k_0)^{-1} \notag
\\
&\quad =
\rho_k^{-1}\big[U_\pm(z,k-1,k_0)+z\alpha_k^*V_\pm(z,k-1,k_0)\big]
\big[\alpha_kU_\pm(z,k-1,k_0)+zV_\pm(z,k-1,k_0)\big]^{-1}\widetilde\rho_k \notag
\\
&\quad = \rho_k^{-1}\big[\Phi_\pm(z,k-1)+z\alpha_k^*\big]
\big[\alpha_k\Phi_\pm(z,k-1)+zI_m\big]^{-1}\widetilde\rho_k. \label{3.139}
\end{align}
For $k$ even, one similarly obtains
\begin{align}
&\Phi_\pm(z,k) = U_\pm(z,k,k_0)V_\pm(z,k,k_0)^{-1} \notag
\\
&\quad =
\rho_k^{-1}\big[\alpha_k^*U_\pm(z,k-1,k_0)+V_\pm(z,k-1,k_0)\big]
\big[U_\pm(z,k-1,k_0)+\alpha_kV_\pm(z,k-1,k_0)\big]^{-1}\widetilde\rho_k \notag
\\
&\quad = \rho_k^{-1}\big[z\alpha_k^*+\Phi_\pm(z,k-1)\big]
\big[zI_m+\alpha_k\Phi_\pm(z,k-1)\big]^{-1}\widetilde\rho_k.
\end{align}
\end{proof}
\begin{remark} \label{r3.18}
$(i)$ In the special case $\alpha=\{\alpha_k\}_{k\in{\mathbb{Z}}}=0$, one obtains
\begin{equation}
M_\pm(z,k) = \pm I_m, \quad \Phi_+(z,k)=0, \quad \Phi_-(z,k)^{-1}=0,
\quad z\in\mathbb{C}, \; k\in{\mathbb{Z}}.
\end{equation}
Thus, strictly speaking, one should always consider $\Phi_-^{-1}$
rather than $\Phi_-$ and hence refer to the Riccati-type equation of
$\Phi_-^{-1}$,
\begin{equation}
z\Phi_-(z,k)^{-1}\rho_k^{-1}\alpha_k^*\Phi_-(z,k-1)^{-1} +
\Phi_-(z,k)^{-1}\rho_k^{-1} - z\widetilde\rho_k^{-1}\Phi_-(z,k-1)^{-1} =
\widetilde\rho_k^{-1}\alpha_k, \quad z\in\mathbb{C}\backslash{\partial\hspace*{.2mm}\mathbb{D}}, \; k\in{\mathbb{Z}},
\label{3.142}
\end{equation}
rather than that of $\Phi_-$, etc. In fact, since $M_-(\cdot,k)$ is an anti-Caratheodory matrix and hence $[M_-(z,k)-I_m]$ is invertible (cf.\ \cite[p.\ 137]{SF70}), we should have introduced the Schur matrix
\begin{equation}
\Phi_-(z,k)^{-1} = [M_-(z,k) + I_m] [M_-(z,k) - I_m]^{-1}, \quad
z\in\mathbb{C}\backslash{\partial\hspace*{.2mm}\mathbb{D}},
\end{equation}
rather than the anti-Schur matrix $\Phi_-(\cdot,k)$. For simplicity of notation, we
will typically avoid this complication with $\Phi_-$ and still invoke
$\Phi_-$ rather than $\Phi_-^{-1}$ whenever confusions are unlikely. \\
$(ii)$ We note that $\Phi_\pm(z,k)^{\pm 1}$, $z\in{\partial\hspace*{.2mm}\mathbb{D}}$, $k\in{\mathbb{Z}}$, have
nontangential limits to ${\partial\hspace*{.2mm}\mathbb{D}}$ Lebesgue almost everywhere. In
particular, the Riccati-type equations \eqref{3.138}, and
\eqref{3.142} extend to ${\partial\hspace*{.2mm}\mathbb{D}}$ Lebesgue almost everywhere.
\end{remark}
The Riccati-type equation \eqref{3.138} for the Schur matrix $\Phi_+$ implies the
norm convergent expansion,
\begin{align}
\Phi_+(z,k)&=\sum_{j=1}^\infty \phi_{+,j}(k) z^j, \quad z\in\mathbb{D},
\; k\in{\mathbb{Z}}, \\
\phi_{+,1}(k)&=-\alpha_{k+1}^*, \notag \\
\phi_{+,2}(k)&=-\rho_{k+1} \alpha_{k+2}^* \widetilde\rho_{k+1}, \label{3.144} \\
\phi_{+,j}(k)&=\sum_{\ell=1}^{j-1}
\rho_{k+1}\phi_{+,j-\ell}(k+1)\widetilde\rho_{k+1}^{-1}\alpha_{k+1}\phi_{+,\ell}(k)
+ \rho_{k+1}\phi_{+,j-1}(k+1)\widetilde\rho_{k+1}^{-1}, \; j\geq 3. \notag
\end{align}
Similarly, the corresponding Riccati-type equation \eqref{3.142} for the Schur matrix
$\Phi_-^{-1}$ implies the norm convergent expansion
\begin{align}
\Phi_-(z,k)^{-1} &=\sum_{j=0}^\infty \phi_{-,j}(k) z^j, \quad
z\in\mathbb{D}, \; k\in{\mathbb{Z}}, \\
\phi_{-,0}(k)&=\alpha_{k}, \notag \\
\phi_{-,1}(k)&=\widetilde\rho_{k}\alpha_{k-1}\rho_{k},
\label{3.146} \\
\phi_{-,j}(k)&=-\sum_{\ell=0}^{j-1}
\phi_{-,j-1-\ell}(k-1)\rho_k^{-1}\alpha_{k}^*\phi_{-,\ell}(k)\rho_k +
\widetilde\rho_k^{-1}\phi_{-,j-1}(k-1)\rho_k, \; j\geq 2. \notag
\end{align}
\section{Weyl--Titchmarsh Theory for CMV Operators on ${\mathbb{Z}}$ with Matrix-valued Verblunsky Coefficients} \label{s4}
In this section we present the basics of Weyl--Titchmarsh
theory for CMV operators on the lattice ${\mathbb{Z}}$ with matrix-valued Verblunsky coefficients. The corresponding case of scalar-valued Verblunsky coefficients was dealt with in detail in \cite{GZ06}.
We start by introducing the ${\mathbb{C}^{m\times m}}$-valued CMV Wronskian of two ${\mathbb{C}^{m\times m}}$-valued sequences $U_j(z,\cdot)$, $j=1,2$,
\begin{align}
W(U_1(1/\overline{z},k),U_2(z,k)) &= \frac{(-1)^{k+1}}{2} \Big[
U_1(1/\overline{z},k)^*U_2(z,k)-({\mathbb{V}}^*U_1(1/\overline{z},\,\cdot\,))(k)^*
({\mathbb{V}}^*U_2(z,\,\cdot\,))(k)\Big], \notag
\\
& \hspace*{7cm} k\in{\mathbb{Z}}, \; z\in{\C\backslash\{0\}}, \label{4.1A}
\end{align}
Next we verify that the Wronskian of any
two solutions of ${\mathbb{U}} U(z,\cdot)=zU(z,\cdot)$ is indeed $k$-independent as expected:
\begin{lemma}
Suppose $U_j(z,\cdot)$ satisfy ${\mathbb{U}} U_j(z,\cdot)=zU_j(z,\cdot)$,
$j=1,2$, where ${\mathbb{U}}$ is understood as a difference expression $($rather
then an operator in $\ell^2({\mathbb{Z}})^{m\times m}$$)$.\ Then the Wronskian in \eqref{4.1A} is independent of $k\in{\mathbb{Z}}$ and the following identities hold:
\begin{align}
W(U_1(1/\overline{z},k),U_2(z,k)) &= \frac{(-1)^{k+1}}{2}
\Big[U_1(1/\overline{z},k)^*U_2(z,k)-V_1(1/\overline{z},k)^*V_2(z,k)\Big] \notag
\\
&= \frac{(-1)^{k+1}}{2}
\begin{pmatrix}
U_1(1/\overline{z},k) \\ V_1(1/\overline{z},k)
\end{pmatrix}^*
\begin{pmatrix}
I_m & 0 \\ 0 & -I_m
\end{pmatrix}
\begin{pmatrix}
U_2(z,k) \\ V_2(z,k)
\end{pmatrix}, \quad k\in{\mathbb{Z}}, \; z\in{\C\backslash\{0\}}, \label{4.2A}
\end{align}
where $V_j(z,\cdot)={\mathbb{V}}^* U_j(z,\cdot)$, $j=1,2$, and
\begin{align}
W(P_+(1/\overline{z},k,k_0),Q_+(z,k,k_0)) &= I_m, \label{4.3A}
\\
W(U_+(1/\overline{z},k,k_0),U_-(z,k,k_0)) &= M_+(z,k_0)-M_-(z,k_0), \quad
k,k_0\in{\mathbb{Z}}, \; z\in{\C\backslash\{0\}}. \label{4.4A}
\end{align}
\end{lemma}
\begin{proof}
First, we note that \eqref{4.2A} is implied by
\eqref{4.1A}. Next, employing Lemma \ref{l3.2},
$U_j$ and $V_j$, $j=1,2$, satisfy the recursion relation
\begin{align}
\binom{U_j(z,k)}{V_j(z,k)} = {\mathbb{T}}(z,k) \binom{U_j(z,k-1)}{V_j(z,k-1)},
\quad j=1,2,\; k\in{\mathbb{Z}},\; z\in{\C\backslash\{0\}},
\end{align}
and hence
\begin{align}
&W(U_1(1/\overline{z},k),U_2(z,k)) = \frac{(-1)^{k+1}}{2}
\begin{pmatrix}
U_1(1/\overline{z},k) \\ V_1(1/\overline{z},k)
\end{pmatrix}^*
\begin{pmatrix}
I_m & 0 \\ 0 & -I_m
\end{pmatrix}
\begin{pmatrix}
U_2(z,k) \\ V_2(z,k)
\end{pmatrix} \notag
\\
&\qquad = \frac{(-1)^{k+1}}{2}
\begin{pmatrix}
U_1(1/\overline{z},k-1) \\ V_1(1/\overline{z},k-1)
\end{pmatrix}^*
{\mathbb{T}}(1/\overline{z},k)^*
\begin{pmatrix}
I_m & 0 \\ 0 & -I_m
\end{pmatrix}
{\mathbb{T}}(z,k)
\begin{pmatrix}
U_2(z,k-1) \\ V_2(z,k-1)
\end{pmatrix} \notag
\\
&\qquad = -\frac{(-1)^{k}}{2}
\begin{pmatrix}
U_1(1/\overline{z},k-1) \\ V_1(1/\overline{z},k-1)
\end{pmatrix}^*
\begin{pmatrix}
-I_m & 0 \\ 0 & I_m
\end{pmatrix}
\begin{pmatrix}
U_2(z,k-1) \\ V_2(z,k-1)
\end{pmatrix}
\\
&\qquad = W(U_1(1/\overline{z},k-1),U_2(z,k-1)), \quad k\in{\mathbb{Z}},\;
z\in{\C\backslash\{0\}}. \notag
\end{align}
Here we used the following identity which is implied by \eqref{3.12}
and \eqref{3.22}
\begin{align}
{\mathbb{T}}(1/\overline{z},k)^*
\begin{pmatrix}
I_m & 0 \\ 0 & -I_m
\end{pmatrix}
{\mathbb{T}}(z,k)
=
\begin{pmatrix}
-I_m & 0 \\ 0 & I_m
\end{pmatrix}, \quad k\in{\mathbb{Z}}, \; z\in{\C\backslash\{0\}}.
\end{align}
Finally, taking $k=k_0$ and utilizing \eqref{3.49}, \eqref{3.50},
\eqref{3.125}, \eqref{3.126}, and \eqref{A.7}, one obtains
\eqref{4.3A} and \eqref{4.4A} from \eqref{4.2A}.
\end{proof}
For notational simplicity we abbreviate the Wronskian of $U_+$ and
$U_-$ by
\begin{equation}
W(z,k_0)=W(U_+(1/\overline{z},k,k_0),U_-(z,k,k_0)).
\end{equation}
Then, using \eqref{3.128}, \eqref{3.131}, and \eqref{4.4A}, one analytically continues $W(z,k_0)$ to $z=0$ and obtains
\begin{align}
W(z,k_0) = M_+(z,k_0)-M_-(z,k_0), \quad k\in{\mathbb{Z}}, \; z\in{\mathbb{C}}.
\label{4.1}
\end{align}
Moreover, one verifies a certain symmetry property of the Wronskian $W(z,k_0)$,
\begin{align}
M_+(z,k_0)W(z,k_0)^{-1}M_-(z,k_0) =
M_-(z,k_0)W(z,k_0)^{-1}M_+(z,k_0), \quad k\in{\mathbb{Z}}, \; z\in\mathbb{C}.
\label{4.1a}
\end{align}
Next we prove an auxiliary lemma that will play a crucial role in our
computation of the resolvent for the CMV operator ${\mathbb{U}}$.
\begin{lemma} \label{l4.1}
Let $k,k_0\in{\mathbb{Z}}$ and $z\in{\C\backslash\{0\}}$. The the following identities hold,
\begin{align}
& P_+(z,k,k_0)Q_+(1/\overline{z},k,k_0)^* +
Q_+(z,k,k_0)P_+(1/\overline{z},k,k_0)^* = 2(-1)^{k+1}I_m, \label{4.2}
\\
& R_+(z,k,k_0)S_+(1/\overline{z},k,k_0)^* +
S_+(z,k,k_0)R_+(1/\overline{z},k,k_0)^* = 2(-1)^{k}I_m, \label{4.3}
\\
& P_+(z,k,k_0)S_+(1/\overline{z},k,k_0)^* +
Q_+(z,k,k_0)R_+(1/\overline{z},k,k_0)^* = 0, \label{4.4}
\\
& R_+(z,k,k_0)Q_+(1/\overline{z},k,k_0)^* +
S_+(z,k,k_0)P_+(1/\overline{z},k,k_0)^* = 0, \label{4.5}
\end{align}
and
\begin{align}
& U_+(z,k,k_0)W(z,k_0)^{-1}U_-(1/\overline{z},k,k_0)^* -
U_-(z,k,k_0)W(z,k_0)^{-1}U_+(1/\overline{z},k,k_0)^* = 2(-1)^{k+1}I_m,
\label{4.6}
\\
& V_+(z,k,k_0)W(z,k_0)^{-1}U_-(1/\overline{z},k,k_0)^* -
V_-(z,k,k_0)W(z,k_0)^{-1}U_+(1/\overline{z},k,k_0)^* = 0. \label{4.7}
\end{align}
\end{lemma}
\begin{proof}
First, we note that for $k=k_0$ equalities \eqref{4.2}--\eqref{4.5}
follow from \eqref{3.49}. Then one uses an induction argument
to prove \eqref{4.2}--\eqref{4.5} for $k\neq k_0$. This involves a
consideration of a number of cases all of which follow the same
pattern. Therefore, we limit out attention to just one of these cases.
Suppose \eqref{4.2}--\eqref{4.5} hold for some $k\in{\mathbb{Z}}$ even.
Then utilizing \eqref{3.21} together with \eqref{3.8} and \eqref{3.9},
one computes
\begin{align}
& P_+(z,k+1,k_0)Q_+(1/\overline{z},k+1,k_0)^* +
Q_+(z,k+1,k_0)P_+(1/\overline{z},k+1,k_0)^* \notag
\\
&\quad = \widetilde\rho_{k+1}^{-1}\alpha_{k+1}
\big[P_+(z,k,k_0)Q_+(1/\overline{z},k,k_0)^* +
Q_+(z,k,k_0)P_+(1/\overline{z},k,k_0)^*\big] \alpha_{k+1}^*\widetilde\rho_{k+1}^{-1}
\notag
\\
&\qquad + \widetilde\rho_{k+1}^{-1} \big[R_+(z,k,k_0)S_+(1/\overline{z},k,k_0)^*
+ S_+(z,k,k_0)R_+(1/\overline{z},k,k_0)^*\big] \widetilde\rho_{k+1}^{-1}
\\
&\qquad + z\widetilde\rho_{k+1}^{-1} \big[R_+(z,k,k_0)Q_+(1/\overline{z},k,k_0)^*
+ S_+(z,k,k_0)P_+(1/\overline{z},k,k_0)^*\big]
\alpha_{k_0}^*\widetilde\rho_{k+1}^{-1} \notag
\\
&\qquad + \widetilde\rho_{k+1}^{-1}\alpha_{k_0}
\big[P_+(z,k,k_0)S_+(1/\overline{z},k,k_0)^* +
Q_+(z,k,k_0)R_+(1/\overline{z},k,k_0)^*\big] \widetilde\rho_{k+1}^{-1}z^{-1} \notag
\\
&\quad = 2(-1)^{k+1}
\big[\widetilde\rho_{k+1}^{-1}\alpha_{k+1}\alpha_{k+1}^*\widetilde\rho_{k+1}^{-1} -
\widetilde\rho_{k+1}^{-2}\big] = 2(-1)^{(k+1)+1}I_m. \notag
\end{align}
Similarly, one checks equalities \eqref{4.3}--\eqref{4.5} at the
point $k+1$. Then inverting the matrix ${\mathbb{T}}(z,k)$ and utilizing
\eqref{3.21} in the form,
\begin{align}
\binom{P_-(z,k-1),k_0}{R_-(z,k-1,k_0)} = {\mathbb{T}}(z,k)^{-1}
\binom{P_-(z,k,k_0)}{R_-(z,k,k_0)},
\end{align}
one verifies \eqref{4.2}--\eqref{4.5} at the point $k-1$. Similarly,
one verifies \eqref{4.2}--\eqref{4.5} at the points $k+1$ and $k-1$
under the assumption of $k$ odd.
Next, using \eqref{3.125}, \eqref{3.126}, \eqref{4.1}, \eqref{4.1a},
\eqref{4.2}, and \eqref{4.5}, one verifies \eqref{4.6} and
\eqref{4.7} as follows:
\begin{align}
& U_+(z,k,k_0)W(z,k_0)^{-1}U_-(1/\overline{z},k,k_0)^* -
U_-(z,k,k_0)W(z,k_0)^{-1}U_+(1/\overline{z},k,k_0)^* \notag
\\
&\quad =
Q_+(z,k,k_0)W(z,k_0)^{-1}\big[M_+(z,k_0)-M_-(z,k_0)\big]P_+(1/\overline{z},k,k_0)^*
\notag
\\
&\qquad +
P_+(z,k,k_0)\big[M_+(z,k_0)-M_-(z,k_0)\big]W(z,k_0)^{-1}Q_+(1/\overline{z},k,k_0)^*
\notag
\\
&\qquad + P_+(z,k,k_0)\big[M_-(z,k_0)W(z,k_0)^{-1}M_+(z,k_0) \notag
\\
&\qquad\quad -
M_+(z,k_0)W(z,k_0)^{-1}M_-(z,k_0)\big]P_+(1/\overline{z},k,k_0)^* \notag
\\
&\quad = Q_+(z,k,k_0)P_+(1/\overline{z},k,k_0)^* +
P_+(z,k,k_0)Q_+(1/\overline{z},k,k_0)^* = 2(-1)^{k+1}I_m,
\\
& V_+(z,k,k_0)W(z,k_0)^{-1}U_-(1/\overline{z},k,k_0)^* -
V_-(z,k,k_0)W(z,k_0)^{-1}U_+(1/\overline{z},k,k_0)^* \notag
\\
&\quad =
S_+(z,k,k_0)W(z,k_0)^{-1}\big[M_+(z,k_0)-M_-(z,k_0)\big]P_+(1/\overline{z},k,k_0)^*
\notag
\\
&\qquad +
R_+(z,k,k_0)\big[M_+(z,k_0)-M_-(z,k_0)\big]W(z,k_0)^{-1}Q_+(1/\overline{z},k,k_0)^*
\notag
\\
&\qquad + R_+(z,k,k_0)\big[M_+(z,k_0)W(z,k_0)^{-1}M_-(z,k_0) \notag
\\
&\qquad\quad -
M_-(z,k_0)W(z,k_0)^{-1}M_+(z,k_0)\big]P_+(1/\overline{z},k,k_0)^* \notag
\\
&\quad = S_+(z,k,k_0)P_+(1/\overline{z},k,k_0)^* +
R_+(z,k,k_0)Q_+(1/\overline{z},k,k_0)^* = 0.
\end{align}
\end{proof}
\begin{lemma} \label{l4.2}
Let $z\in{\mathbb{C}}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\})$ and fix $k_0\in{\mathbb{Z}}$. Then the
resolvent $({\mathbb{U}}-zI)^{-1}$ of the unitary CMV operator ${\mathbb{U}}$ on
$\ltm{{\mathbb{Z}}}$ is given in terms of its matrix representation in the
standard basis of $\ltm{{\mathbb{Z}}}$ by
\begin{align}
({\mathbb{U}}-zI)^{-1}(k,k') = \frac{1}{2z}
\begin{cases}
U_-(z,k,k_0)W(z,k_0)^{-1}U_+(1/\overline{z},k',k_0)^*, & k < k' \text{ or
} k = k' \text{ odd},
\\
U_+(z,k,k_0)W(z,k_0)^{-1}U_-(1/\overline{z},k',k_0)^*, & k > k' \text{ or
} k = k' \text{ even},
\end{cases} \label{4.13}
\\
\quad k,k' \in{\mathbb{Z}}. \notag
\end{align}
Moreover, since $0\in{\mathbb{C}}\backslash\sigma({\mathbb{U}})$, \eqref{4.13}
analytically extends to $z=0$.
In particular, one obtains for any $z\in{\mathbb{C}}\backslash{\partial\hspace*{.2mm}\mathbb{D}}$,
\begin{align}
&({\mathbb{U}}-zI)^{-1}(k,k) \notag
\\
&\quad = \frac{1}{2z}
\begin{cases}
[I_m+M_-(z,k)] W(z,k)^{-1} [I_m-M_+(z,k)], & k \text{ odd},\\
[I_m-M_+(z,k)] W(z,k)^{-1} [I_m+M_-(z,k)], & k \text{ even},
\end{cases} \label{4.14}
\\
&({\mathbb{U}}-zI)^{-1}(k-1,k-1) \notag
\\
&\quad = \frac{1}{2z}
\begin{cases}
\rho_{k}^{-1}[a_{k}^*-b_{k}^*M_+(z,k)] W(z,k)^{-1}
[a_{k}+M_-(z,k)b_{k}]\rho_{k}^{-1}, & k \text{ odd},\\
\widetilde\rho_{k}^{-1}[a_{k}+b_{k}M_-(z,k)] W(z,k)^{-1}
[a_{k}^*-M_+(z,k)b_{k}^*]\widetilde\rho_{k}^{-1}, & k \text{
even},
\end{cases} \label{4.15}
\\
&({\mathbb{U}}-zI)^{-1}(k-1,k) \notag
\\
&\quad = \frac{-1}{2z}
\begin{cases}
\rho_{k}^{-1} [a_{k}^*-b_{k}^*M_-(z,k)] W(z,k)^{-1}
[I_m-M_+(z,k)], & k \text{ odd}, \\
\widetilde\rho_{k}^{-1} [a_{k}+b_{k}M_-(z,k)] W(z,k)^{-1}
[I_m+M_+(z,k)], & k \text{ even},
\end{cases} \label{4.16}
\\
&({\mathbb{U}}-zI)^{-1}(k,k-1) \notag
\\
&\quad = \frac{-1}{2z}
\begin{cases}
[I_m+M_+(z,k)] W(z,k)^{-1} [a_{k}+M_-(z,k)b_{k}]
\rho_{k}^{-1}, & k \text{ odd},
\\
[I_m-M_+(z,k)] W(z,k)^{-1} [a_{k}^*-M_-(z,k)b_{k}^*]
\widetilde\rho_{k}^{-1}, & k \text{ even}.
\end{cases} \label{4.17}
\end{align}
\end{lemma}
\begin{proof}
Let
\begin{align}
G(z,k,k',k_0) =
\begin{cases}
U_-(z,k,k_0)W(z,k_0)^{-1}U_+(1/\overline{z},k',k_0)^*, & k < k' \text{ or
} k = k' \text{ odd},
\\
U_+(z,k,k_0)W(z,k_0)^{-1}U_-(1/\overline{z},k',k_0)^*, & k > k' \text{ or
} k = k' \text{ even},
\end{cases}
\\
\quad k,k' \in{\mathbb{Z}}. \notag
\end{align}
Then \eqref{4.13} is equivalent to
\begin{align}
&({\mathbb{U}}-zI)G(z,\cdot,k',k_0) = 2z\Delta_{k'},\quad k',k_0\in{\mathbb{Z}}. \label{4.19}
\end{align}
First, assume $k'$ to be odd. Then,
\begin{align}
(({\mathbb{U}}-zI)G(z,\cdot,k',k_0))(\ell) =
(({\mathbb{V}}{\mathbb{W}}-zI)G&(z,\cdot,k',k_0))(\ell) = 0, \quad \ell \in {\mathbb{Z}}\backslash
\{k',k'+1\}, \label{4.20}
\end{align}
and by \eqref{4.6}, \eqref{4.7},
\begin{align}
&\binom{(({\mathbb{U}}-zI)G(z,\cdot,k',k_0))(k')}
{(({\mathbb{U}}-zI)G(z,\cdot,k',k_0))(k'+1)} =
\binom{(({\mathbb{V}}{\mathbb{W}}-zI)G(z,\cdot,k',k_0))(k')}
{(({\mathbb{V}}{\mathbb{W}}-zI)G(z,\cdot,k',k_0))(k'+1)} \notag
\\
&\quad = \Theta_{k'+1}
\binom{zV_-(z,k',k_0)W(z,k_0)^{-1}U_+(1/\overline{z},k',k_0)^*}
{zV_+(z,k'+1,k_0)W(z,k_0)^{-1}U_-(1/\overline{z},k',k_0)^*} -
z\binom{G(z,k',k',k_0)}{G(z,k'+1,k',k_0)} \notag
\\
&\quad = z\Theta_{k'+1}
\binom{V_+(z,k',k_0)W(z,k_0)^{-1}U_-(1/\overline{z},k',k_0)^*}
{V_+(z,k'+1,k_0)W(z,k_0)^{-1}U_-(1/\overline{z},k',k_0)^*} -
z\binom{G(z,k',k',k_0)}{G(z,k'+1,k',k_0)} \label{4.21}
\\
&\quad = z\binom{U_+(z,k',k_0)W(z,k_0)^{-1}U_-(1/\overline{z},k',k_0)^*}
{U_+(z,k'+1,k_0)W(z,k_0)^{-1}U_-(1/\overline{z},k',k_0)^*} -
z\binom{G(z,k',k',k_0)}{G(z,k'+1,k',k_0)} \notag
\\
&\quad = z\binom{2(-1)^{k'+1}I_m}{0} = \binom{2zI_m}{0}. \notag
\end{align}
Thus, for $k'$ odd, \eqref{4.19} is a consequence of \eqref{4.20} and
\eqref{4.21}.
Next, assume $k'$ to be even. Then,
\begin{align}
(({\mathbb{U}}-zI)G(z,\cdot,k',k_0))(\ell) = (({\mathbb{V}}{\mathbb{W}}-zI)G(z,\cdot,k',k_0))(\ell)
= 0, \quad \ell \in {\mathbb{Z}}\backslash \{k'-1,k'\}, \label{4.22}
\end{align}
and by \eqref{4.6}, \eqref{4.7},
\begin{align}
&\binom{(({\mathbb{U}}-zI)G(z,\cdot,k',k_0))(k'-1)}
{(({\mathbb{U}}-zI)G(z,\cdot,k',k_0))(k')} =
\binom{(({\mathbb{V}}{\mathbb{W}}-zI)G(z,\cdot,k',k_0))(k'-1)}
{(({\mathbb{V}}{\mathbb{W}}-zI)G(z,\cdot,k',k_0))(k')} \notag
\\
& \quad = \Theta_{k'}
\binom{zV_-(z,k'-1,k_0)W(z,k_0)^{-1}U_+(1/\overline{z},k',k_0)^*}
{zV_+(z,k',k_0)W(z,k_0)^{-1}U_-(1/\overline{z},k',k_0)^*} -
z\binom{G(z,k'-1,k',k_0)}{G(z,k',k',k_0)} \notag
\\
& \quad = z\Theta_{k'}
\binom{V_-(z,k'-1,k_0)W(z,k_0)^{-1}U_+(1/\overline{z},k',k_0)^*}
{V_-(z,k',k_0)W(z,k_0)^{-1}U_+(1/\overline{z},k',k_0)^*} -
z\binom{G(z,k'-1,k',k_0)}{G(z,k',k',k_0)} \label{4.23}
\\
& \quad =
z\binom{U_-(z,k'-1,k_0)W(z,k_0)^{-1}U_+(1/\overline{z},k',k_0)^*}
{U_-(z,k',k_0)W(z,k_0)^{-1}U_+(1/\overline{z},k',k_0)^*} -
z\binom{G(z,k'-1,k',k_0)}{G(z,k',k',k_0)} \notag
\\
&\quad = z\binom{0}{2(-1)^{k'}I_m} = \binom{0}{2zI_m}. \notag
\end{align}
Thus, for $k'$ even, \eqref{4.19} follows from \eqref{4.22} and
\eqref{4.23}, and hence one obtains \eqref{4.13}.
Finally, using \eqref{Table} and \eqref{3.125} one verifies the identities
\begin{align}
U_\pm(z,k,k) &=
\begin{cases}
z[I_m+M_\pm(z,k)], & k \text{ odd},\\
-I_m+M_\pm(z,k), & k \text{ even},
\end{cases} \label{4.24}
\\
U_\pm(z,k-1,k) &=
\begin{cases}
-z\rho_{k}^{-1}[a_{k}^*-b_{k}^*M_\pm(z,k)], & k \text{ odd},\\
\widetilde\rho_{k}^{-1}[a_{k}+b_{k}M_\pm(z,k)], & k \text{ even}.
\end{cases} \label{4.25}
\end{align}
Inserting \eqref{4.24} and \eqref{4.25} into \eqref{4.13} and
utilizing the fact that (anti-)Caratheodory matrices satisfy
$M_\pm(1/\overline{z},k)^* = -M_\pm(z,k)$, $k\in{\mathbb{Z}}$, $z\in\mathbb{C}$, one obtains
\eqref{4.14}--\eqref{4.17}.
\end{proof}
Next, we briefly turn to Weyl--Titchmarsh theory for CMV operators
with matrix-valued Verblunsky coefficients on ${\mathbb{Z}}$. We denote by
$d\Omega(\cdot,k)$, $k\in{\mathbb{Z}}$, the $2m \times 2m$ matrix-valued
measure,
\begin{align}
d\Omega(\zeta,k) &= d
\begin{pmatrix}
\Omega_{0,0}(\zeta,k) & \Omega_{0,1}(\zeta,k)
\\
\Omega_{1,0}(\zeta,k) & \Omega_{1,1}(\zeta,k)
\end{pmatrix} \notag
\\ &= d
\begin{pmatrix}
\Delta_{k-1}^*E_{{\mathbb{U}}}(\zeta)\Delta_{k-1} & \Delta_{k-1}^*E_{{\mathbb{U}}}(\zeta)\Delta_{k}
\\
\Delta_{k}^*E_{{\mathbb{U}}}(\zeta)\Delta_{k-1} & \Delta_{k}^*E_{{\mathbb{U}}}(\zeta)\Delta_{k}
\end{pmatrix}, \quad \zeta \in{\partial\hspace*{.2mm}\mathbb{D}}, \label{4.26}
\end{align}
where $E_{{\mathbb{U}}}(\cdot)$ denotes the family of spectral projections
of the unitary CMV operator ${\mathbb{U}}$ on $\ltm{{\mathbb{Z}}}$,
\begin{equation}
{\mathbb{U}}=\oint_{\partial\hspace*{.2mm}\mathbb{D}} dE_{{\mathbb{U}}}(\zeta)\,\zeta.
\end{equation}
We also introduce the $2m \times 2m$ matrix-valued function
${\mathcal M}(\cdot,k)$, $k\in{\mathbb{Z}}$, by
\begin{align}
&{\mathcal M}(z,k) = \begin{pmatrix} M_{0,0}(z,k) & M_{0,1}(z,k) \\
M_{1,0}(z,k) & M_{1,1}(z,k) \end{pmatrix} \notag
\\ &\quad =
\begin{pmatrix}
\Delta_{k-1}^*({\mathbb{U}}+zI)({\mathbb{U}}-zI)^{-1}\Delta_{k-1}
&\Delta_{k-1}^*({\mathbb{U}}+zI)({\mathbb{U}}-zI)^{-1}\Delta_{k}
\\
\Delta_{k}^*({\mathbb{U}}+zI)({\mathbb{U}}-zI)^{-1}\Delta_{k-1} &
\Delta_{k}^*({\mathbb{U}}+zI)({\mathbb{U}}-zI)^{-1}\Delta_{k}
\end{pmatrix} \label{4.28}
\\ &\quad =
\oint_{\partial\hspace*{.2mm}\mathbb{D}} d\Omega(\zeta,k)\, \frac{\zeta+z}{\zeta-z}, \quad
z\in{\mathbb{C}}\backslash{\partial\hspace*{.2mm}\mathbb{D}}. \notag
\end{align}
We note that
\begin{align}
M_{0,0}(\cdot,k+1) = M_{1,1}(\cdot,k), \quad k\in{\mathbb{Z}} \label{4.29}
\end{align}
and
\begin{align}
\begin{split}
M_{1,1}(z,k) &= \Delta_{k}^*({\mathbb{U}}+zI)({\mathbb{U}}-zI)^{-1}\Delta_{k} \label{4.30}
\\ & = \oint_{\partial\hspace*{.2mm}\mathbb{D}} d\Omega_{1,1}(\zeta,k) \,
\frac{\zeta+z}{\zeta-z}, \quad z\in{\mathbb{C}}\backslash{\partial\hspace*{.2mm}\mathbb{D}},\; k\in{\mathbb{Z}},
\end{split}
\end{align}
where
\begin{equation}
d\Omega_{1,1}(\zeta,k)=d\Delta_{k}^*E_{{\mathbb{U}}}(\zeta)\Delta_{k}, \quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}}.
\label{4.32}
\end{equation}
Thus, $M_{0,0}|_{\mathbb{D}}$ and $M_{1,1}|_{\mathbb{D}}$ are $m\times m$
Caratheodory matrices. Moreover, by \eqref{4.30} one
infers that
\begin{equation}
M_{1,1}(0,k)=I_m, \quad k\in{\mathbb{Z}}. \label{4.33}
\end{equation}
\begin{lemma}
Let $z\in{\mathbb{C}}\backslash{\partial\hspace*{.2mm}\mathbb{D}}$. Then the functions
$M_{\ell,\ell'}(\cdot,k)$, $\ell,\ell'=0,1$, and $M_\pm(\cdot,k)$,
$k\in{\mathbb{Z}}$, satisfy the following relations
\begin{align}
&M_{0,0}(z,k) = I_m +
\begin{cases}
\rho_{k}^{-1}[a_{k}^*-b_{k}^*M_+(z,k)] W(z,k)^{-1}
[a_{k}+M_-(z,k)b_{k}]\rho_{k}^{-1}, & k \text{ odd},\\
\widetilde\rho_{k}^{-1}[a_{k}+b_{k}M_-(z,k)] W(z,k)^{-1}
[a_{k}^*-M_+(z,k)b_{k}^*] \widetilde\rho_{k}^{-1}, & k \text{ even},
\end{cases} \label{4.34}
\\
&M_{1,1}(z,k) = I_m +
\begin{cases}
[I_m+M_-(z,k)] W(z,k)^{-1} [I_m-M_+(z,k)], & k \text{ odd},\\
[I_m-M_+(z,k)] W(z,k)^{-1} [I_m+M_-(z,k)], & k \text{ even},
\end{cases} \label{4.35}
\\
&M_{0,1}(z,k) = -
\begin{cases}
\rho_{k}^{-1} [a_{k}^*-b_{k}^*M_-(z,k)] W(z,k)^{-1}
[I_m-M_+(z,k)], & k \text{ odd}, \\
\widetilde\rho_{k}^{-1} [a_{k}+b_{k}M_-(z,k)] W(z,k)^{-1} [I_m+M_+(z,k)],
& k \text{ even},
\end{cases} \label{4.36}
\\
&M_{1,0}(z,k) = -
\begin{cases}
[I_m+M_+(z,k)] W(z,k)^{-1} [a_{k}+M_-(z,k)b_{k}] \rho_{k}^{-1}, & k
\text{ odd},
\\
[I_m-M_+(z,k)] W(z,k)^{-1} [a_{k}^*-M_-(z,k)b_{k}^*]
\widetilde\rho_{k}^{-1}, & k \text{ even},
\end{cases} \label{4.37}
\end{align}
where $a_k=I_m+\alpha_k$ and $b_k=I_m-\alpha_k$, $k\in{\mathbb{Z}}$.
\end{lemma}
\begin{proof}
The result is a consequence of Lemma \ref{l4.2} since by \eqref{4.28} one
has
\begin{align}
M_{\ell,\ell'}(z,k) &= \Delta_{k-1+\ell}^*(I +
2z({\mathbb{U}}-zI)^{-1})\Delta_{k-1+\ell'} \notag
\\
&= I_m\delta_{\ell,\ell'} + ({\mathbb{U}}-zI)^{-1}(k-1+\ell,k-1+\ell').
\end{align}
\end{proof}
Finally, introducing the $m\times m$ matrix-valued functions
$\Phi_{1,1}(\cdot,k)$, $k\in{\mathbb{Z}}$, by
\begin{align} \label{4.39}
\Phi_{1,1}(z,k) &= [M_{1,1}(z,k)-I_m][M_{1,1}(z,k)+I_m]^{-1} \notag
\\
&= I_m - 2[M_{1,1}(z,k)+I_m]^{-1}, \quad z\in\mathbb{C}\backslash{\partial\hspace*{.2mm}\mathbb{D}},
\end{align}
then,
\begin{align}
M_{1,1}(z,k) &= [I_m+\Phi_{1,1}(z,k)][I_m-\Phi_{1,1}(z,k)]^{-1} \notag
\\
&=2[I_m-\Phi_{1,1}(z,k)]^{-1} - I_m, \quad z\in\mathbb{C}\backslash{\partial\hspace*{.2mm}\mathbb{D}}.
\label{4.40}
\end{align}
\begin{lemma}
The ${\mathbb{C}^{m\times m}}$-valued function $\Phi_{1,1}|_{\mathbb{D}}$ is a Schur matrix and $\Phi_{1,1}$
is related to $\Phi_\pm$ by
\begin{equation}
\Phi_{1,1}(z,k) =
\begin{cases}
\Phi_-(z,k)^{-1}\Phi_+(z,k), & k\text{ odd},
\\
\Phi_+(z,k)\Phi_-(z,k)^{-1}, & k\text{ even},
\end{cases}
\quad z\in\mathbb{C}\backslash{\partial\hspace*{.2mm}\mathbb{D}},\; k\in{\mathbb{Z}}. \label{4.41}
\end{equation}
\end{lemma}
\begin{proof}
Suppose $k$ is odd. Then\eqref{4.1}, \eqref{4.1a},
and \eqref{4.35} imply that
\begin{align}
M_{1,1}(z,k)+I_m &= [I_m+M_-(z,k)] W(z,k)^{-1} [I_m-M_+(z,k)] \notag
\\
&\quad + [M_+(z,k)-M_-(z,k)] W(z,k)^{-1} +
W(z,k)^{-1} [M_+(z,k)-M_-(z,k)] \notag
\\
&= [I_m+M_+(z,k)] W(z,k)^{-1} [I_m-M_-(z,k)]. \label{4.42}
\end{align}
Using \eqref{3.133}, \eqref{4.1}, \eqref{4.39}, and \eqref{4.42},
one computes
\begin{align}
\Phi_{1,1}(z,k) &= I_m-2[I_m-M_-(z,k)]^{-1}W(z,k)[I_m+M_+(z,k)]^{-1}
\notag
\\
&=[I_m-M_-(z,k)]^{-1} \big[[I_m-M_-(z,k)][I_m+M_+(z,k)]-2W(z,k)\big]
[I_m+M_+(z,k)]^{-1} \notag
\\
&=[I_m-M_-(z,k)]^{-1}[I_m+M_-(z,k)]
[I_m-M_+(z,k)][I_m+M_+(z,k)]^{-1}
\\
&=\Phi_-(z,k)^{-1}\Phi_+(z,k), \quad z\in\mathbb{C}\backslash{\partial\hspace*{.2mm}\mathbb{D}},\;
k\in{\mathbb{Z}}. \notag
\end{align}
The result for $k$ even is proved similarly.
\end{proof}
Next we introduce a sequence of $\mathbb{C}^{m\times 2m}$-valued
Laurent polynomials $\{P(z,k,k_0)\}_{k\in{\mathbb{Z}}}$ by
\begin{align} \label{4.44}
P(z,k,k_0) &= \big(P_0(z,k,k_0),P_1(z,k,k_0)\big) \notag
\\
&=
\begin{cases}
\big(P_+(z,k,k_0),Q_+(z,k,k_0)\big)
\begin{pmatrix}
\frac{1}{2z}\rho_{k_0} & \frac{1}{2z}a_{k_0}^*
\\
-\frac{1}{2z}\rho_{k_0} & \frac{1}{2z}b_{k_0}^*
\end{pmatrix}, & k_0 \text{ odd},
\\[4mm]
\big(P_+(z,k,k_0),Q_+(z,k,k_0)\big)
\begin{pmatrix}
\frac{1}{2}\widetilde\rho_{k_0} & \frac{1}{2}a_{k_0}
\\ \frac{1}{2}\widetilde\rho_{k_0} & -\frac{1}{2}b_{k_0}
\end{pmatrix}, & k_0 \text{ even}.
\end{cases}
\end{align}
Then it is easy to see that $P_j(z,\cdot,k_0)$, $j=0,1$ are linear
combinations of $P_+(z,\cdot,k_0)$ and $Q_+(z,\cdot,k_0)$, and hence
satisfy ${\mathbb{U}} P_j(z,\cdot,k_0) = zP_j(z,\cdot,k_0)$, $j=0,1$.
Moreover, \eqref{Table} and \eqref{4.44} imply that
\begin{align}
\begin{split}
P(z,k_0-1,k_0) &= (P_0(z,k_0-1,k_0),P_1(z,k_0-1,k_0)) = (I_m,0),
\\
P(z,k_0,k_0) &= (P_0(z,k_0,k_0),P_1(z,k_0,k_0)) = (0,I_m),
\end{split} \label{4.45}
\end{align}
and hence any solution $U(z,\cdot)$ of ${\mathbb{U}} U(z,\cdot) = zU(z,\cdot)$
can be expressed as
\begin{align} \label{4.46}
U(z,\cdot) = P_0(z,\cdot,k_0)U(z,k_0-1)+P_1(z,\cdot,k_0)U(z,k_0).
\end{align}
Our next goal is to show that the Laurent polynomials
$\{P(z,k,k_0)^*\}_{k\in{\mathbb{Z}}}$ form complete orthonormal system in
$\Ltm{}$. To do that we first prove an auxiliary result analogous to
Lemma \ref{l3.5}.
\begin{lemma} \label{l4.5}
Suppose $\{F(\cdot,k)\}_{k\in{\mathbb{Z}}}$ is a sequence of ${\mathbb{C}^{m\times m}}$-valued
functions of bounded variation with $F(1,k)=0$ for all $k\in{\mathbb{Z}}$
that satisfies
\begin{align}
({\mathbb{U}} F(\zeta,\cdot))(k) = \int_{A_{\zeta}} dF(\zeta',k) \, \zeta',
\quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\; k\in{\mathbb{Z}},
\end{align}
where ${\mathbb{U}}$ are understood in the sense of difference expressions
rather than difference operators on $\ltm{{\mathbb{Z}}}$. Then, $F(\cdot,k)$ also
satisfies
\begin{align}
F(\zeta,k) = \int_{A_{\zeta}}P_0(\zeta',k,k_0)\,dF(\zeta',k_0-1) +
\int_{A_{\zeta}}P_1(\zeta',k,k_0)\,dF(\zeta',k_0), \quad
\zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\; k, k_0\in{\mathbb{Z}}. \label{4.48}
\end{align}
\end{lemma}
\begin{proof}
Let $\{G(\cdot,k,k_0)\}_{k\in{\mathbb{Z}}}$ denote the sequence of
${\mathbb{C}^{m\times m}}$-valued functions,
\begin{align}
G(\zeta,k,k_0) = \int_{A_{\zeta}}P_0(\zeta',k,k_0)\,dF(\zeta',k_0-1)
+ \int_{A_{\zeta}}P_1(\zeta',k,k_0)\,dF(\zeta',k_0), \quad
\zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\; k, k_0\in{\mathbb{Z}}. \label{4.49}
\end{align}
Then it suffices to prove that $F(\zeta,k) = G(\zeta,k,k_0)$,
$\zeta\in{\partial\hspace*{.2mm}\mathbb{D}}$, $k,k_0\in{\mathbb{Z}}$.
First, we note that \eqref{4.45} and \eqref{4.49} imply that
\begin{align}
\begin{split}
G(\zeta,k_0-1,k_0) &=
\int_{A_{\zeta}}dF(\zeta',k_0-1)=F(\zeta,k_0-1),
\\
G(\zeta,k_0,k_0) &= \int_{A_{\zeta}}dF(\zeta',k_0)=F(\zeta,k_0),
\quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\; k_0\in{\mathbb{Z}},
\end{split}
\end{align}
and
\begin{align}
({\mathbb{U}} G(\zeta,\cdot,k_0))(k) &= \int_{A_{\zeta}} ({\mathbb{U}}
P_0(\zeta',\cdot,k_0))(k)\,dF(\zeta',k_0-1) + \int_{A_{\zeta}} ({\mathbb{U}}
P_1(\zeta',\cdot,k_0))(k)\,dF(\zeta',k_0) \notag
\\
&= \int_{A_{\zeta}} dG(\zeta',k,k_0) \, \zeta', \quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\;
k,k_0\in{\mathbb{Z}}.
\end{align}
Next, defining $K(\zeta,k,k_0) = F(\zeta,k)-G(\zeta,k,k_0)$,
$\zeta\in{\partial\hspace*{.2mm}\mathbb{D}}$, $k,k_0\in{\mathbb{Z}}$, one obtains
\begin{align*}
\begin{split}
&K(\zeta,k_0-1,k_0) = K(\zeta,k_0,k_0) = 0,
\\
&({\mathbb{U}} K(\zeta,\cdot,k_0))(k) = \int_{A_{\zeta}} dK(\zeta',k,k_0) \,
\zeta' , \quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\; k,k_0\in{\mathbb{Z}},
\end{split}
\end{align*}
or equivalently,
\begin{align}
\begin{split}
&K(\zeta,k_0-1,k_0) = K(\zeta,k_0,k_0) = 0, \label{4.52}
\\
&({\mathbb{U}} K(\zeta,\cdot,k_0))(k) = (\L \, K(\cdot,k,k_0))(\zeta), \quad
\zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\; k,k_0\in{\mathbb{Z}},
\end{split}
\end{align}
where $\L$ denotes the boundedly invertible operator on ${\mathbb{C}^{m\times m}}$-valued
functions $K$ of bounded variation defined by
\begin{align}
(\L \, K)(\zeta) = \int_{A_{\zeta}} dK(\zeta')\, \zeta', \quad
(\L^{-1} K)(\zeta) = \int_{A_{\zeta}} dK(\zeta') \, {\zeta'}^{-1}.
\end{align}
Finally, since, $\L$ commutes with all constant $m\times m$
matrices, one can repeat the proof of Lemma \ref{l3.2} with $z$
replaced by $\L$ and using \eqref{4.46} obtain that
\eqref{4.52} has the unique solution $K(\zeta,k,k_0)=0$,
$\zeta\in{\partial\hspace*{.2mm}\mathbb{D}}$, $k,k_0\in{\mathbb{Z}}$, and hence, $F(\zeta,k) =
G(\zeta,k,k_0)$, $\zeta\in{\partial\hspace*{.2mm}\mathbb{D}}$, $k,k_0\in{\mathbb{Z}}$.
\end{proof}
\begin{lemma} \label{l4.6}
Let $k_0\in{\mathbb{Z}}$. Then the set of $\mathbb{C}^{2m\times m}$-valued Laurent polynomials
$\{P(\cdot,k,k_0)^*\}_{k\in{\mathbb{Z}}}$ forms a complete orthonormal system
on ${\partial\hspace*{.2mm}\mathbb{D}}$ with respect to $\mathbb{C}^{2m\times2m}$-valued measure
$d\Omega(\cdot,k_0)$. Explicitly, $P(\cdot,k,k_0)$, $k\in{\mathbb{Z}}$, satisfy,
\begin{align}
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} P(\zeta,k,k_0)\,d\Omega(\zeta,k_0)\,P(\zeta,k',k_0)^* &=
\delta_{k,k'}I_m, \quad k,k'\in{\mathbb{Z}} \label{4.54}
\end{align}
and the collection of $\mathbb{C}^{2m}$-valued Laurent polynomials
\begin{align}
\left\{
\begin{pmatrix}
(P(\cdot,k,k_0))_{1,j}\\\vdots\\(P(\cdot,k,k_0))_{2m,j}
\end{pmatrix}\right\}_{ j=1,\dots,2m, \, k\in{\mathbb{Z}}}
\end{align}
form complete systems in $\Ltm{}$.
\end{lemma}
\begin{proof}
Fix an integer $k'\in{\mathbb{Z}}$ and let $\{F(\cdot,k,k')\}_{k\in{\mathbb{Z}}}$ denote
the ${\mathbb{C}^{m\times m}}$-valued sequences of functions of bounded
variation defined by
\begin{align}
F(\zeta,k,k') = \Delta_{k}^* E_{{\mathbb{U}}}(\zeta)\Delta_{k'}, \quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}},
\; k\in{\mathbb{Z}}. \label{4.55}
\end{align}
Then,
\begin{align}
({\mathbb{U}} F(\zeta,\cdot,k'))(k) &= ({\mathbb{U}} E_{{\mathbb{U}}}(\zeta)\Delta_{k'})(k) =
\left(\int_{A_{\zeta}} dE_{{\mathbb{U}}}(\zeta')\, \zeta' \Delta_{k'}\right)(k)
\\ &=
\int_{A_{\zeta}} d\big(\Delta_k^* E_{{\mathbb{U}}}(\zeta')\Delta_{k'}\big) \, \zeta'
= \int_{A_{\zeta}} dF(\zeta',k,k') \, \zeta', \quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}},\;
k\in{\mathbb{Z}}, \notag
\end{align}
and hence \eqref{4.48} in Lemma \ref{l4.5} implies that
\begin{align} \label{4.57}
dF(\zeta,k,k') = P_0(\zeta,k,k_0)\,dF(\zeta,k_0-1,k') +
P_1(\zeta,k,k_0)\,dF(\zeta,k_0,k'), \quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}}, \; k\in{\mathbb{Z}},
\end{align}
or equivalently,
\begin{align} \label{4.58}
dF(\zeta,k',k) = dF(\zeta,k',k)^* =
dF(\zeta,k',k_0-1)\,P_0(\zeta,k,k_0)^* +
dF(\zeta,k',k_0)\,P_1(\zeta,k,k_0)^*,& \notag
\\
\quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}}, \; k\in{\mathbb{Z}}.&
\end{align}
In particular, taking $k'=k_0-1$ and $k'=k_0$, one obtains from
\eqref{4.58},
\begin{align}
dF(\zeta,k_0-1,k) &= dF(\zeta,k_0-1,k_0-1)\,P_0(\zeta,k,k_0)^* +
dF(\zeta,k_0-1,k_0)\,P_1(\zeta,k,k_0)^*, \label{4.59}
\\
dF(\zeta,k_0,k) &= dF(\zeta,k_0,k_0-1)\,P_0(\zeta,k,k_0)^* +
dF(\zeta,k_0,k_0)\,P_1(\zeta,k,k_0)^*,\quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}}, \; k\in{\mathbb{Z}}.
\notag
\end{align}
Next, setting $k=k'$ in \eqref{4.59} and plugging it into
\eqref{4.57}, one obtains
\begin{align}
dF(\zeta,k,k') = \sum_{\ell,\ell'=0}^1 P_\ell(\zeta,k,k_0)
\,dF(\zeta,k_0-1+\ell,k_0-1+\ell')\, P_{\ell'}(\zeta,k',k_0)^*,
\quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}}, \; k,k'\in{\mathbb{Z}}. \label{4.60}
\end{align}
Integrating \eqref{4.60} over the unit circle ${\partial\hspace*{.2mm}\mathbb{D}}$ and observing
that by \eqref{4.26} and \eqref{4.55}
$dF(\zeta,k_0-1+\ell,k_0-1+\ell') = d\Omega_{\ell,\ell'}(z,k_0)$,
$\ell,\ell'=0,1$, one obtains
\begin{align}
\delta_{k,k'}I_m = \oint_{\partial\hspace*{.2mm}\mathbb{D}}\sum_{\ell,\ell'=0}^1 P_\ell(\zeta,k,k_0)
\,d\Omega_{\ell,\ell'}(\zeta,k_0)\, P_{\ell'}(\zeta,k',k_0)^*, \quad
\zeta\in{\partial\hspace*{.2mm}\mathbb{D}}, \; k,k'\in{\mathbb{Z}},
\end{align}
which is equivalent to \eqref{4.54}.
To prove completeness of $\{P(\cdot,k,k_0)^*\}_{k\in{\mathbb{Z}}}$ we first note
the fact,
\begin{align}
{\text{\rm span}}\{P(z,k,k_0)^*\}_{k\in{\mathbb{Z}}} &= {\text{\rm span}} \left\{
\begin{pmatrix} z^k I_m \\ z^{k-1} I_m\end{pmatrix},
\begin{pmatrix} z^{k-1} I_m \\ z^k I_m\end{pmatrix},
\begin{pmatrix} I_m \\ 0\end{pmatrix},
\begin{pmatrix} 0 \\ I_m\end{pmatrix} \right\}_{k\in{\mathbb{Z}}} \notag \\
&= {\text{\rm span}}\left\{\begin{pmatrix} z^k I_m\\ 0 \end{pmatrix},
\begin{pmatrix} 0 \\ z^k I_m\end{pmatrix} \right\}_{k\in{\mathbb{Z}}}.
\end{align}
This is a consequence of investigating the leading-order coefficients
of $P_+(z,k,k_0)$ and $Q_+(z,k,k_0)$ (cf.\ Remark \ref{r2.4}) and
\eqref{4.44}). Thus, it suffices to prove that
$\Big\{\Big(\begin{smallmatrix}\zeta^k I_m \\ 0\end{smallmatrix}\Big),
\Big(\begin{smallmatrix}0 \\ \zeta^k I_m
\end{smallmatrix}\Big)\Big\}_{k\in{\mathbb{Z}}}$ is a complete system
with respect to $d\Omega(\cdot,k_0)$.
Let $F=\Big(\begin{smallmatrix} F_0 \\
F_1 \end{smallmatrix}\Big)\in\Ltm{}$ and suppose $F$ is orthogonal
to all columns of $\Big(\begin{smallmatrix} \zeta^k I_m\\ 0
\end{smallmatrix}\Big)$ and $\Big(\begin{smallmatrix} 0 \\
\zeta^k I_m\end{smallmatrix}\Big)$ for all $k\in{\mathbb{Z}}$, that is,
\begin{equation}
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \begin{pmatrix} \zeta^k I_m \\ 0\end{pmatrix}^*\,
d\Omega (\zeta,k_0)\, F(\zeta) = \oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \zeta^{-k} \,
[d\Omega_{0,0}(\zeta,k_0)\,F_0(\zeta)
+d\Omega_{0,1}(\zeta,k_0)\,F_1(\zeta)] =
\begin{pmatrix}0\\\vdots\\0\end{pmatrix}\in\mathbb{C}^{2m} \label{4.63}
\end{equation}
and
\begin{equation}
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \begin{pmatrix} 0 \\ \zeta^k I_m \end{pmatrix}^*\,
d\Omega (\zeta,k_0)\, F(\zeta) =\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \zeta^{-k} \,
[d\Omega_{1,0}(\zeta,k_0)\,F_0(\zeta) +
d\Omega_{1,1}(\zeta,k_0)\,F_1(\zeta)] =
\begin{pmatrix}0\\\vdots\\0\end{pmatrix}\in\mathbb{C}^{2m} \label{4.64}
\end{equation}
for all $k\in{\mathbb{Z}}$. Note that for a scalar complex-valued measure
$d\omega$ equalities $\oint d\omega(\zeta)\,\zeta^n=0$, $n\in{\mathbb{Z}}$, imply $\oint
d\Re(\omega(\zeta))\,\zeta^n=\oint d\Im(\omega(\zeta))\,\zeta^n = 0$, and hence
\cite[p.\ 24]{Du83}) implies that $d\omega=0$. Applying this
argument to $d(\Omega_{0,0}F_0 + \Omega_{0,1}F_1)_\ell$ and
$d(\Omega_{1,0}F_0 + \Omega_{1,1}F_1)_\ell$, $\ell=1,\dots,2m$, one
obtains
\begin{align}
d\Omega_{0,0}F_0 + d\Omega_{0,1}F_1 &=
\begin{pmatrix}0\\\vdots\\0\end{pmatrix}\in\mathbb{C}^{2m}, \label{4.65}
\\
d\Omega_{1,0}F_0 + d\Omega_{1,1}F_1 &=
\begin{pmatrix}0\\\vdots\\0\end{pmatrix}\in\mathbb{C}^{2m}. \label{4.66}
\end{align}
Multiplying \eqref{4.65} by $F_0^*$ on the left and \eqref{4.66} by
$F_1^*$ on the left and adding the results then yields
\begin{align}
\|F\|_{\Ltm{}}^2=\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} F(\zeta)^* \,d\Omega(\zeta,k_0)\,
F(\zeta) = 0.
\end{align}
\end{proof}
\begin{corollary}
The full-lattice CMV operator ${\mathbb{U}}$ is unitarily equivalent to the
operator of multiplication by $\zeta$ on $\Ltm{}$ for any
$k_0\in{\mathbb{Z}}$. In particular,
\begin{align}
& \sigma({\mathbb{U}}) = \text{\rm{supp}} \, (d\Omega(\cdot,k_0)), \quad k_0\in{\mathbb{Z}}.
\end{align}
\end{corollary}
\begin{proof}
Consider the linear map
$\dot{\mathcal U}\colon\ltzm{{\mathbb{Z}}}\to\Ltm{}$ from the space of compactly
supported sequences $\ltzm{{\mathbb{Z}}}$ to the set of $\mathbb{C}^{2m}$-valued
Laurent polynomials defined by
\begin{equation}
(\dot {\mathcal U} F)(z) = \sum_{k=-\infty}^{\infty} P(1/\overline{z},k,k_0)^*
F(k), \quad F\in \ltzm{{\mathbb{Z}}}.
\end{equation}
Using \eqref{4.54} one shows that $\widehat F(\zeta) = (\dot {\mathcal U}
F)(\zeta)$, $F\in \ltzm{{\mathbb{Z}}}$ has the property
\begin{align}
\|\widehat F\|^2_{\Ltm{}} &= \oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \widehat
F(\zeta)^*d\Omega(\zeta,k_0)\,\widehat F(\zeta)
\\ &=
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \sum_{k=-\infty}^\infty F(k)^* P(\zeta,k,k_0)
\,d\Omega_\pm(\zeta,k_0)\!\! \sum_{k'=-\infty}^{\infty}
P_\pm(\zeta,k',k_0)^*F(k') \notag
\\ &=
\sum_{k,k'=-\infty}^{\infty} F(k)^* \left(\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} P(\zeta,k,k_0)
\,d\Omega(\zeta,k_0)\, P(\zeta,k',k_0)^*\right)F(k') \notag
\\ &=
\sum_{k=-\infty}^{\infty} F(k)^*F(k) = \|F\|^2_{\ltm{{\mathbb{Z}}}}. \label{4.70}
\end{align}
Since $\ltzm{{\mathbb{Z}}}$ is dense in $\ltm{{\mathbb{Z}}}$, $\dot {\mathcal U}$ extends by
continuity to a bounded linear operator ${\mathcal U}\colon \ltm{{\mathbb{Z}}} \to
\Ltm{}$, and the identity
\begin{align} \label{4.71}
({\mathcal U}({\mathbb{U}} F))(\zeta) &= \sum_{k=-\infty}^{\infty} P(\zeta,k,k_0)^* ({\mathbb{U}}
F)(k) = \sum_{k=-\infty}^{\infty} ({\mathbb{U}}^* P(\zeta,\cdot,k_0))(k)^*
F(k)
\\ &=
\sum_{k=-\infty}^{\infty} (\zeta^{-1} P(\zeta,k,k_0))^* F(k) =
\zeta({\mathcal U} F)(\zeta), \quad F\in\ltm{{\mathbb{Z}}}, \notag
\end{align}
holds. The range of the operator ${\mathcal U}$ is all of $\Ltm{}$ since the
$\mathbb{C}^{2m\times m}$-valued Laurent polynomials $\{P(\cdot,k,k_0)^*\}_{k\in{\mathbb{Z}}}$
are complete with respect to $d\Omega(\cdot,k_0)$. Hence the inverse
operator ${\mathcal U}^{-1}$ exists on $\Ltm{}$ and is given by
\begin{equation}
({\mathcal U}^{-1}\widehat F)(k) = \oint_{{\partial\hspace*{.2mm}\mathbb{D}}}
P(\zeta,k,k_0)\,d\Omega(\zeta,k_0)\,\widehat F(\zeta), \quad \widehat F\in
\Ltm{},
\end{equation}
which together with \eqref{4.70} implies that ${\mathcal U}$ is unitary. In
addition, \eqref{4.71} shows that the full-lattice unitary
operator ${\mathbb{U}}$ on $\ltm{{\mathbb{Z}}}$ is unitarily equivalent to the
operators of multiplication by $\zeta$ on $\Ltm{}$,
\begin{align}
({\mathcal U} {\mathbb{U}} {\mathcal U}^{-1} \widehat F)(\zeta) = \zeta\widehat F(\zeta), \quad \widehat
F\in \Ltm{}.
\end{align}
\end{proof}
\section{Borg--Marchenko-type Uniqueness Results for CMV Operators with
Matrix-valued Verblunsky Coefficients} \label{s5}
In this section we prove (local and global) Borg--Marchenko-type uniqueness results for
CMV operators with matrix-valued Verblunsky coefficients on half-lattices and on
the full lattice ${\mathbb{Z}}$. We freely use the notation established in Sections \ref{s3},
\ref{s4}, and Appendix \ref{sA}.
We start with uniqueness results on half-lattices.
\begin{theorem} \label{t5.1}
Assume Hypothesis \ref{h3.1} and let $k_0\in{\mathbb{Z}}$, $N\in{\mathbb{N}}$. Then for
the right half-lattice problem the following sets of data $(i)$--$(v)$ are
equivalent:
\begin{align}
(i) &\quad \big\{\alpha_{k_0+k}\big\}_{k=1}^N.
\\
(ii) &\quad \bigg\{\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}d\Omega_+(\zeta,k_0)\, \zeta^k \bigg\}_{k=1}^N.
\\
(iii) &\quad \big\{m_{+,k}(k_0)\big\}_{k=1}^N, \;\text{ where
$m_{+,k}(k_0)$, $k\geq 0$, are the Taylor coefficients of
$m_+(z,k_0)$} \notag\\ &\hspace*{5mm} \text{at $z=0$, that is, }\;
m_+(z,k_0) = \sum\nolimits_{k=0}^\infty m_{+,k}(k_0)z^k, \; z\in\mathbb{D}.
\\
(iv) &\quad \big\{M_{+,k}(k_0)\big\}_{k=1}^N, \;\text{ where
$M_{+,k}(k_0)$, $k\geq 0$, are the Taylor coefficients of
$M_+(z,k_0)$} \notag\\ &\hspace*{5mm} \text{at $z=0$, that is, }\;
M_+(z,k_0) = \sum\nolimits_{k=0}^\infty M_{+,k}(k_0)z^k, \; z\in\mathbb{D}.
\\
(v) &\quad \big\{\phi_{+,k}(k_0)\big\}_{k=1}^N, \;\text{ where
$\phi_{+,k}(k_0)$, $k\geq 0$, are the Taylor coefficients of
$\Phi_+(z,k_0)$} \notag\\ &\hspace*{5mm} \text{at $z=0$, that is, }\;
\Phi_+(z,k_0) = \sum\nolimits_{k=0}^\infty \phi_{+,k}(k_0)z^k, \;
z\in\mathbb{D}.
\end{align}
Similarly, for the left half-lattice problem, the following sets of
data $(vi)$--$(x)$ are equivalent:
\begin{align}
(vi) &\quad \big\{\alpha_{k_0-k}\big\}_{k=0}^{N-1}.
\\
(vii) &\quad \bigg\{\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}d\Omega_-(\zeta,k_0)\, \zeta^k \bigg\}_{k=1}^N.
\\
(viii) &\quad \big\{m_{-,k}(k_0)\big\}_{k=1}^{N}, \;\text{ where
$m_{-,k}(k_0)$, $k\geq 0$, are the Taylor coefficients of
$m_-(z,k_0)$} \notag\\ &\hspace*{5mm} \text{at $z=0$, that is, }\;
m_-(z,k_0) = \sum\nolimits_{k=0}^\infty m_{-,k}(k_0)z^k.
\\
(ix) &\quad \big\{M_{-,k}(k_0)\big\}_{k=0}^{N-1}, \;\text{ where
$M_{-,k}(k_0)$, $k\geq 0$, are the Taylor coefficients of
$M_-(z,k_0)$} \notag\\ &\hspace*{5mm} \text{at $z=0$, that is, }\;
M_-(z,k_0) = \sum\nolimits_{k=0}^\infty M_{-,k}(k_0)z^k.
\\
(x) &\quad \big\{\phi_{-,k}(k_0)\big\}_{k=0}^{N-1}, \;\text{ where
$\phi_{-,k}(k_0)$, $k\geq 0$, are the Taylor coefficients of
$\Phi_-(z,k_0)^{-1}$} \notag\\ &\hspace*{5mm} \text{at $z=0$, that is, }\;
\Phi_-(z,k_0)^{-1} = \sum\nolimits_{k=0}^\infty \phi_{-,k}(k_0)z^k.
\end{align}
\end{theorem}
\begin{proof}
$(i)\Rightarrow(ii)$ and $(vi)\Rightarrow(vii)$: First, utilizing
relations \eqref{3.37} and \eqref{3.40} with the initial conditions
\eqref{3.49} and \eqref{3.50}, one constructs $\{P_\pm(z,k_0\pm
k,k_0)\}_{k=1}^{N}$ and $\{R_\pm(z,k_0\pm k,k_0)\big\}_{k=1}^{N}$.
We note that the Laurent polynomials
\begin{align}
&\begin{cases}%
z^{-1}P_+(z,k_0+k,k_0), \; R_-(z,k_0-k,k_0), & k_0 \text{ odd},
\\[1mm]
R_+(z,k_0+k,k_0), \; z^{-1}P_-(z,k_0-k,k_0), & k_0 \text{ even},
\end{cases} \label{5.15}
\intertext{are linear combinations of }
&\begin{cases}%
I_m,z^{-1}I_m,zI_m,z^{-2}I_m,z^2I_m,\dots,z^{(k-1)/2}I_m,z^{-(k+1)/2}I_m,
& k \text{ odd},
\\[1mm]
I_m,z^{-1}I_m,zI_m,z^{-2}I_m,z^2I_m,\dots,z^{-k/2}I_m,z^{k/2}I_m, & k
\text{ even},
\end{cases} \label{5.16}
\intertext{and}
&\begin{cases}%
R_+(z,k_0+k,k_0), \; P_-(z,k_0-k,k_0), & k_0 \text{ odd},
\\[1mm]
P_+(z,k_0+k,k_0), \; R_-(z,k_0-k,k_0), & k_0 \text{ even},
\end{cases} \label{5.17}
\intertext{are linear combinations of }
&\begin{cases}%
I_m,zI_m,z^{-1}I_m,z^2I_m,z^{-2}I_m,\dots,z^{-(k-1)/2}I_m,z^{(k+1)/2}I_m,
& k \text{ odd},
\\[1mm]
I_m,zI_m,z^{-1}I_m,z^2I_m,z^{-2}I_m,\dots,z^{k/2}I_m,z^{-k/2}I_m, & k
\text{ even}.
\end{cases} \label{5.18}
\end{align}
Moreover, the last elements of the sequences in \eqref{5.16} and
\eqref{5.18} represent the leading-order terms of the Laurent polynomials in
\eqref{5.15} and \eqref{5.17}, respectively, and the corresponding
leading-order coefficients are invertible $m \times m$ matrices (cf.\ Remark \ref{r2.4}).
Next, assume $k_0$ and $k$ to be odd. Then utilizing \eqref{5.17}
and \eqref{5.18} one finds $m\times m$ matrices $C_{\pm,j}$ and
$D_{\pm,j}$, $0\leq j\leq k$, such that
\begin{align}
z^{-(k-1)/2}I_m &= \sum_{j=0}^{k} C_{+,j}\, R_+(z,k_0+j, k_0),
\quad z^{(k+1)/2}I_m = \sum_{j=0}^{k} D_{+,j}\, R_+(z,k_0+j,k_0),
\\
z^{-(k-1)/2}I_m &= \sum_{j=0}^{k} C_{-,j}\, P_-(z,k_0-j, k_0),
\quad z^{(k+1)/2}I_m = \sum_{j=0}^{k} D_{-,j}\, P_-(z,k_0-j,k_0),
\end{align}
and, using \eqref{3.56} and \eqref{3.57}, computes
\begin{align}
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}d\Omega_\pm(\zeta,k_0)\, \zeta^k =
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} \big(\zeta^{(k+1)/2}I_m\big) \, d\Omega_\pm(\zeta,k_0)
\, \big(\zeta^{-(k-1)/2}I_m\big)^*
= \sum_{j=0}^k D_{\pm,j}\,C_{\pm,j}^*.
\end{align}
The other cases of $k_0$ and $k$ follow similarly.
$(ii)\Rightarrow(i)$ and $(vii)\Rightarrow(vi)$: Since
$d\Omega_\pm(\cdot,k_0)$ are nonnegative normalized measures, one has
\begin{align}
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}d\Omega_\pm(\zeta,k_0)\,\zeta^{-k} =
\left(\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}d\Omega_\pm(\zeta,k_0)\,\zeta^{k}\right)^*
\, \text{ and } \, \oint_{{\partial\hspace*{.2mm}\mathbb{D}}}d\Omega_\pm(\zeta,k_0) = I_m, \label{5.22}
\end{align}
that is, the knowledge of positive moments imply the knowledge of
negative ones. Applying Corollary \ref{c3.8} one constructs the matrix-valued
orthonormal Laurent polynomials $\{P_\pm(\zeta,k_0\pm k,k_0)\}_{k=1}^{N}$ and
$\{R_\pm(\zeta,k_0\pm k,k_0)\big\}_{k=1}^{N}$. Subsequently applying
Theorem \ref{t3.9}, in particular, formulas \eqref{3.83} and
\eqref{3.84}, one obtains the coefficients $(i)$ and $(vi)$.
$(ii)\Leftrightarrow(iii)$ and $(vii)\Leftrightarrow(viii)$: These
follow from \eqref{3.111} and \eqref{5.22},
\begin{align}
m_\pm(z,k_0) &= \pm
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}d\Omega_{\pm}(\zeta,k_0)\,\frac{\zeta+z}{\zeta-z} = \pm
I_m \pm 2\sum_{k=1}^{\infty}
z^k\left(\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}d\Omega_\pm(\zeta,k_0)\,\zeta^{k}\right)^*, \quad
z\in\mathbb{D}.
\end{align}
$(iii)\Leftrightarrow(iv)$: This is implied by \eqref{3.127}.
$(iv)\Leftrightarrow(v)$: This is a consequence of \eqref{3.133} and
\eqref{3.134}, together with the facts: For $|z|$ sufficiently small,
$\|M_+(z,k_0)-I_m\|_{\mathbb{C}^{m\times m}}<1$ by \eqref{3.128}, and
$\|\Phi_+(z,k_0)\|_{\mathbb{C}^{m\times m}}<1$ by \eqref{3.133a}. Hence,
\begin{align}
M_+(z,k_0) &= [I_m-\Phi_+(z,k_0)]^{-1}[I_m+\Phi_+(z,k_0)] \notag \\
& \hspace{-1.5mm}
\underset{z\to 0}{=} [I_m+\Phi_+(z,k_0)] \sum_{k=0}^\infty \Phi_+(z,k_0)^k,
\\
\Phi_+(z,k_0) &= \big[2^{-1}[M_+(z,k_0)-I_m]\big]
\big[I_m+2^{-1}[M_+(z,k_0)-I_m]\big]^{-1}
\notag \\
& \hspace{-1.5mm}
\underset{z\to 0}{=} -\sum_{k=1}^\infty 2^{-k}[I_m-M_+(z,k_0)]^k.
\end{align}
$(ix)\Leftrightarrow(x)$: This is implied by \eqref{3.131}, \eqref{3.133}, \eqref{3.134},
and the fact that, for $|z|$ sufficiently small, $\|\Phi_-(z,k_0)^{-1}\|_{{\mathbb{C}^{m\times m}}}<1$ by
\eqref{3.7} and \eqref{3.133a}. Hence,
\begin{align}
M_-(z,k_0) &= [\Phi_-(z,k_0)^{-1}-I_m]^{-1}[\Phi_-(z,k_0)^{-1}+I_m]
\notag
\\
& \hspace{-1.5mm}
\underset{z\to 0}{=} -[\Phi_-(z,k_0)^{-1}+I_m]\sum_{k=0}^\infty \Phi_-(z,k_0)^{-k},
\\
\Phi_-(z,k_0)^{-1} &=
[M_-(z,k_0)+I_m][M_-(z,k_0)-M_-(0,k_0)+M_-(0,k_0)-I_m]^{-1} \notag
\\ &=
\big[[M_-(z,k_0)+I_m][M_-(0,k_0)-I_m]^{-1}\big] \notag
\\
&\quad \times \big[[M_-(z,k_0)-M_-(0,k_0)][M_-(0,k_0)-I_m]^{-1}+I_m\big]^{-1}
\\ & \hspace{-1.5mm}
\underset{z\to 0}{=}
\big[[M_-(z,k_0)+I_m][M_-(0,k_0)-I_m]^{-1}\big] \notag
\\
&\quad \times \sum_{k=0}^\infty
\big[[M_-(z,k_0)-M_-(0,k_0)][I_m-M_-(0,k_0)]^{-1}\big]^k. \notag
\end{align}
$(viii)\Leftrightarrow(x)$: This follows because \eqref{3.111a}, \eqref{3.135}, and
the fact that $\|\Phi_-(z,k_0)^{-1}\|_{{\mathbb{C}^{m\times m}}}\leq1$, $z\in\mathbb{D}$, together imply that
\begin{align}
m_-(z,k_0) &= [z\Phi_-(z,k_0)^{-1}+I_m]^{-1}[z\Phi_-(z,k_0)^{-1}-I_m]
\notag \\
& \hspace{-1.5mm}
\underset{z\to 0}{=} [z\Phi_-(z,k_0)^{-1}-I_m] \sum_{k=0}^\infty
\big[-z\Phi_-(z,k_0)^{-1}\big]^k,
\\
z\Phi_-(z,k_0)^{-1} &= [I_m+m_-(z,k_0)][I_m-m_-(z,k_0)]^{-1} \notag
\\
&= 2^{-1} [I_m+m_-(z,k_0)]\big[I_m- 2^{-1}[I_m+m_-(z,k_0)]\big]^{-1}
\\
& \hspace{-1.5mm}
\underset{z\to 0}{=} \sum_{k=1}^\infty 2^{-k}[I_m+m_-(z,k_0)]^k. \notag
\qedhere
\end{align}
\end{proof}
Next, we restate Theorem \ref{t5.1}:
\begin{theorem} \label{t5.2}
Assume Hypothesis \ref{h3.1} for two sequences $\alpha^{(1)}$,
$\alpha^{(2)}$ and let $k_0\in{\mathbb{Z}}$, $N\in{\mathbb{N}}$. Then for the right
half-lattice problems associated with $\alpha^{(1)}$ and $\alpha^{(2)}$
the following items $(i)$--$(iv)$ are equivalent:
\begin{align}
(i) &\quad \alpha_k^{(1)} = \alpha_k^{(2)}, \quad k_0+1\leq k\leq k_0+N.
\\
(ii) &\quad m_+^{(1)}(z,k_0)-m_+^{(2)}(z,k_0) \underset{z\to 0}{=} o(z^N).
\\
(iii) &\quad M_+^{(1)}(z,k_0)-M_+^{(2)}(z,k_0) \underset{z\to 0}{=} o(z^N).
\\
(iv) &\quad \Phi_+^{(1)}(z,k_0)-\Phi_+^{(2)}(z,k_0) \underset{z\to 0}{=} o(z^N).
\end{align}
Similarly, for the left half-lattice problems associated with
$\alpha^{(1)}$ and $\alpha^{(2)}$, the following items $(v)$--$(viii)$ are equivalent:
\begin{align}
(v) &\quad \alpha_k^{(1)} = \alpha_k^{(2)}, \quad k_0-N+1\leq k\leq k_0.
\\
(vi) &\quad m_-^{(1)}(z,k_0)-m_-^{(2)}(z,k_0) \underset{z\to 0}{=} o(z^N).
\\
(vii) &\quad M_-^{(1)}(z,k_0)-M_-^{(2)}(z,k_0) \underset{z\to 0}{=} o(z^{N-1}).
\\
(viii) &\quad \Phi_-^{(1)}(z,k_0)^{-1}-\Phi_-^{(2)}(z,k_0)^{-1} \underset{z\to 0}{=}
o(z^{N-1}).
\end{align}
\end{theorem}
\begin{proof}
This is an immediate consequence of Theorem \ref{t5.1}.
\end{proof}
Finally, we turn to CMV operators on ${\mathbb{Z}}$ and start with two auxiliary results
that play a role in the proofs of analogous
Borg--Marchenko-type uniqueness results for CMV operators on ${\mathbb{Z}}$.
\begin{lemma} \label{l5.4}
Let $A,B,C,D$ denote some $m\times m$ matrices. Suppose that
$A\neq0$, $B$ is invertible, and $A,B,C,D$ satisfy
\begin{align}
\big[2\sqrt{\norm{A}\norm{D}}+\norm{C}\big]\|B^{-1}\|<1. \label{5.45}
\end{align}
Then the matrix-valued Riccati-type equation
\begin{align}
XAX + BX + XC + D = 0, \quad
\norm{X}<\frac{1-\norm{C}\|B^{-1}\|}{2\norm{A}\|B^{-1}\|}, \label{5.46}
\end{align}
has a unique solution $X\in{\mathbb{C}^{m\times m}}$ given by
\begin{align}
X=\lim_{n\to\infty}X_n \,\text{ with }\, \norm{X}\leq
\frac{1-\|C\|\,\|B^{-1}\|}{2\|A\|\,\|B^{-1}\|} -
\sqrt{\left(\frac{1-\|C\|\,\|B^{-1}\|}{2\|A\|\,\|B^{-1}\|}\right)^2
- \frac{\|D\|}{\|A\|}}, \label{5.47}
\end{align}
where
\begin{equation}
X_0=0, \quad X_n=F(X_{n-1}), \; n\in{\mathbb{N}}, \, \text{ and } \, F(X)=-B^{-1}XAX -
B^{-1}XC - B^{-1}D.
\end{equation}
A similar result also holds if $A\neq0$, $C$ is invertible, and
$A,B,C,D$ satisfying
\begin{align}
\big[2\sqrt{\norm{A}\norm{D}}+\norm{B}\big]\|C^{-1}\|<1. \label{5.45a}
\end{align}
In this case, the matrix-valued Riccati-type equation
\begin{align}
XAX + BX + XC + D = 0, \quad
\norm{X}<\frac{1-\norm{B}\|C^{-1}\|}{2\norm{A}\|C^{-1}\|},
\label{5.46a}
\end{align}
has a unique solution $X\in{\mathbb{C}^{m\times m}}$ given by
\begin{align}
X=\lim_{n\to\infty}X_n \,\text{ with }\, \norm{X}\leq
\frac{1-\|B\|\,\|C^{-1}\|}{2\|A\|\,\|C^{-1}\|} -
\sqrt{\left(\frac{1-\|B\|\,\|C^{-1}\|}{2\|A\|\,\|C^{-1}\|}\right)^2
- \frac{\|D\|}{\|A\|}}, \label{5.47a}
\end{align}
where
\begin{equation}
X_0=0, \quad X_n=G(X_{n-1}), \; n\in{\mathbb{N}}, \, \text{ and } \, G(X)=-XAXC^{-1} -
BXC^{-1} - DC^{-1}.
\end{equation}
\end{lemma}
\begin{proof}
Since $B$ is invertible, the equation for $X$ in
\eqref{5.46} is equivalent to $F(X)=X$. Therefore, it suffices to
show that $F(\cdot)$ is a strict contraction on some closed ball of
radius $\lambda$ centered at the origin, $B_\lambda=\{X\in{\mathbb{C}^{m\times m}} \;|\;
\norm{X}\leq\lambda\}$, and that $F(\cdot)$ preserves $B_\lambda$, that is
$\norm{F(X)}\leq\lambda$ whenever $\norm{X}\leq\lambda$.
First, we check that for any
$\lambda<\frac{1-\norm{C}\|B^{-1}\|}{2\norm{A}\|B^{-1}\|}$, the map
$F(\cdot)$ is a strict contraction on $B_\lambda$. Let $X,Y\in B_\lambda$,
then
\begin{align}
\norm{F(X)-F(Y)} &\leq \big[\norm{A}\|B^{-1}\|\norm{X} +
\norm{A}\|B^{-1}\|\norm{Y} + \norm{C}\|B^{-1}\|\big] \norm{X-Y}
\\
&\leq \big[2\lambda\norm{A}\|B^{-1}\|+\norm{C}\|B^{-1}\|\big] \norm{X-Y},
\quad 2\lambda\norm{A}\|B^{-1}\|+\norm{C}\|B^{-1}\|<1. \notag
\end{align}
Next, we check that $F(\cdot)$ preserves $B_\lambda$ for any $\lambda$
satisfying
\begin{align}
\frac{1-\|C\|\,\|B^{-1}\|}{2\|A\|\,\|B^{-1}\|} -
\sqrt{\left(\frac{1-\|C\|\,\|B^{-1}\|}{2\|A\|\,\|B^{-1}\|}\right)^2
- \frac{\|D\|}{\|A\|}}
\leq\lambda<\frac{1-\|C\|\,\|B^{-1}\|}{2\|A\|\,\|B^{-1}\|}. \label{5.49}
\end{align}
Let $X\in B_\lambda$, then by \eqref{5.49}
\begin{align}
\norm{F(X)}\leq \norm{A}\|B^{-1}\|\lambda^2 + \norm{C}\|B^{-1}\|\lambda +
\norm{D}\|B^{-1}\|\leq\lambda.
\end{align}
Thus, Banach's contraction mapping principle implies that
$F(\cdot)$ has a unique fixed point $X$ for which \eqref{5.46} and
\eqref{5.47} hold.
The second part of the Lemma is proved similarly.
\end{proof}
\begin{corollary} \label{c5.4}
Let $A_j$, $B_j$, $C_j$, $D_j$, $j=1,2$, denote some $m\times m$
matrices. Suppose that either $B_1$ and $B_2$ are invertible and
\begin{align}
0<\norm{A_j},\,\|B^{-1}_j\| \leq a, \quad \norm{C_j},\,\norm{D_j}
\leq b, \quad j=1,2, \label{5.54a}
\end{align}
or $C_1$ and $C_2$ are invertible and
\begin{align}
0<\norm{A_j},\,\|C^{-1}_j\| \leq a, \quad \norm{B_j},\,\norm{D_j}
\leq b, \quad j=1,2, \label{5.55a}
\end{align}
for some $a,b>0$ satisfying $2ab(1+2a^2)\leq1$. Then there exist
unique solutions $X_j$, $j=1,2$, of the matrix-valued Riccati-type
equations
\begin{align}
X_jA_jX_j + B_jX_j + X_jC_j + D_j = 0, \quad
\norm{X_j}<\frac{1-ab}{2a^2}, \quad j=1,2, \label{5.56a}
\end{align}
and the following estimate holds
\begin{align}
\norm{X_1-X_2} \leq \lambda(a,b)\big[\norm{A_1-A_2} + \norm{B_1-B_2} +
\norm{C_1-C_2} + \norm{D_1-D_2}\big], \label{5.57a}
\end{align}
where $\lambda(a,b)$ is given by
\begin{align}
\lambda(a,b) = \frac{\max\left\{a, \frac{2a^2b}{1-ab},
a^2b+\frac{2a^3b^2}{1-ab}+\frac{4a^5b^2}{(1-ab)^2},
\frac{4a^3b^2}{(1-ab)^2}\right\}}
{(1-ab)-\frac{4a^3b}{1-ab}}
> 0. \label{5.58a}
\end{align}
\end{corollary}
\begin{proof}
Suppose $B_j$, $j=1,2$, are invertible and note that
$b\leq1/(2a(1+2a^2))$ implies
\begin{align}
\Big(2\sqrt{\norm{A_j}\norm{D_j}}+\norm{C_j}\Big)\|B^{-1}_j\| \leq
(2\sqrt{ab}+b)a \leq \frac{2a\sqrt{2+4a^2}+1}{2(1+2a^2)} <
\frac{2a(2a+\frac{1}{2a})+1}{2(1+2a^2)} = 1
\end{align}
and
\begin{align}
\frac{1-\|C_j\|\,\|B^{-1}_j\|}{2\|A_j\|\,\|B^{-1}_j\|} -
\sqrt{\left( \frac{1-\|C_j\|\,\|B^{-1}_j\|}{2\|A_j\|\,\|B^{-1}_j\|}
\right)^2 - \frac{\|D_j\|}{\|A_j\|}} \leq \frac{2ab}{1-ab} <
\frac{1-ab}{2a^2} \leq
\frac{1-\norm{C_j}\|B^{-1}_j\|}{2\norm{A_j}\|B^{-1}_j\|}.
\end{align}
Then Lemma \ref{l5.4} implies that the matrix-valued
Riccati-type equations in \eqref{5.56a} have unique solutions $X_j$
satisfying $\norm{X_j}\leq\frac{2ab}{1-ab}$, $j=1,2$ and
$X_j=F_j(X_j)$, where $F_j(X)=-{B_j}^{-1}XA_jX - B^{-1}_jXC_j -
B^{-1}_jD_j$, $j=1,2$. Hence, one computes
\begin{align} \label{5.61a}
\norm{X_1-X_2} &= \norm{F_1(X_1)-F_2(X_2)} \notag
\\
&\leq \norm{B_1^{-1}X_1A_1X_1-B_2^{-1}X_2A_2X_2} +
\norm{B^{-1}_1X_1C_1-B^{-1}_2X_2C_2} +
\norm{B^{-1}_1D_1-B^{-1}_2D_2} \notag
\\
&\leq \big[\norm{A_1}\|B^{-1}_2\|\norm{X_1} +
\norm{A_2}\|B^{-1}_2\|\norm{X_2} +
\|B^{-1}_2\|\norm{C_1}\big]\norm{X_1-X_2} \notag
\\
&\quad + \|B^{-1}_2\|\norm{X_1}\norm{X_2}\norm{A_1-A_2} +
\|B^{-1}_2\|\norm{X_2}\norm{C_1-C_2} + \|B^{-1}_2\|\norm{D_1-D_2}
\notag
\\
&\quad + \big[\norm{A_1}\norm{X_1}^2 + \norm{C_1}\norm{X_1} +
\norm{D_1}\big]\|B^{-1}_1\|\,\|B^{-1}_2\|\norm{B_1-B_2} \notag
\\
&\leq \bigg(\frac{4a^3b}{1-ab}+ab\bigg)\norm{X_1-X_2} +
\bigg(\frac{4a^5b^2}{(1-ab)^2}+\frac{2a^3b^2}{1-ab}+a^2b\bigg)
\norm{B_1-B_2}
\\
&\quad + \frac{4a^3b^2}{(1-ab)^2}\norm{A_1-A_2} +
\frac{2a^2b}{1-ab}\norm{C_1-C_2} + a\norm{D_1-D_2}. \notag
\end{align}
Finally, utilizing $b\leq1/(2a(1+2a^2))$, one verifies that
\begin{align}
1-\bigg(\frac{4a^3b}{1-ab}+ab\bigg) = 1-\frac{4a^3b+ab(1-ab)}{1-ab}
> 1-\frac{ab(1+4a^2)}{1-ab} \geq 1-\frac{1+4a^2}{2(1+2a^2)-1} = 0,
\label{5.62a}
\end{align}
and hence \eqref{5.57a} and \eqref{5.58a} follow from
\eqref{5.61a}, and \eqref{5.62a}.
The case of $C_j$ being invertible, $j=1,2$, is proved analogously.
\end{proof}
Given these preliminaries, we introduce the following notation for
the diagonal and for the neighboring off-diagonal entries of the
Green's matrix of ${\mathbb{U}}$ (i.e., the discrete integral kernel of $({\mathbb{U}}-zI)^{-1}$),
\begin{align}
g(z,k) &= ({\mathbb{U}}-Iz)^{-1}(k,k), \label{5.62b}
\\
h(z,k) &=
\begin{cases}
({\mathbb{U}}-Iz)^{-1}(k-1,k), & k \text{ odd}, \\
({\mathbb{U}}-Iz)^{-1}(k,k-1), & k \text{ even},
\end{cases}\quad k\in{\mathbb{Z}},\; z\in\mathbb{D}. \label{5.62c}
\end{align}
Then the subsequent uniqueness results hold for the full-lattice CMV
operator ${\mathbb{U}}$:
\begin{theorem} \label{t5.4}
Assume Hypothesis \ref{h3.1} and let $k_0\in{\mathbb{Z}}$. Then any of the
following two sets of data
\begin{enumerate}[$(i)$]
\item $g(z,k_0)$ and $h(z,k_0)$ for all $z$ in some open $($nonempty$)$
neighborhood of the origin under the assumption that $h(0,k_0)$ is invertible;
\item $g(z,k_0-1)$ and $g(z,k_0)$ for all $z$ in some open $($nonempty$)$ neighborhood
of the origin and $\alpha_{k_0}$ under the assumption $\alpha_{k_0}$ is invertible;
\end{enumerate}
uniquely determine the matrix-valued Verblunsky coefficients $\{\alpha_k\}_{k\in{\mathbb{Z}}}$, and
hence the full-lattice CMV operator ${\mathbb{U}}$.
\end{theorem}
\begin{proof}
{\it Case} $(i)$. First, we note that \eqref{3.18} implies that
\begin{align}
g(0,k_0) &= ({\mathbb{U}}^{-1})_{k_0,k_0} = ({\mathbb{U}}^*)_{k_0,k_0} =
({\mathbb{U}}_{k_0,k_0})^* =
\begin{cases}
-\alpha_{k_0}\alpha_{k_0+1}^*, & k_0 \text{ odd},\\
-\alpha_{k_0+1}^*\alpha_{k_0}, & k_0 \text{ even},
\end{cases}
\\
h(0,k_0) &=
\begin{cases}
({\mathbb{U}}^{-1})_{k_0-1,k_0} = ({\mathbb{U}}_{k_0,k_0-1})^* =
-\rho_{k_0}\alpha_{k_0+1}^*, & k_0 \text{ odd},
\\
({\mathbb{U}}^{-1})_{k_0,k_0-1} = ({\mathbb{U}}_{k_0-1,k_0})^* =
-\alpha_{k_0+1}^*\widetilde\rho_{k_0}, & k_0 \text{ even}.
\end{cases}
\end{align}
Since $h(0,k_0)$ is invertible, one can solve the above equalities
for $\rho_{k_0}$ and $\alpha_{k_0}$,
\begin{align}
&g(0,k_0)h(0,k_0)^{-1} = \alpha_{k_0}\rho_{k_0}^{-1}, \quad k_0 \text{
odd},
\\
&h(0,k_0)^{-1}g(0,k_0) = \widetilde\rho_{k_0}^{-1}\alpha_{k_0} =
\alpha_{k_0}\rho_{k_0}^{-1}, \quad k_0 \text{ even},
\end{align}
implying
\begin{align}
&\rho_{k_0} =
\begin{cases}
\big[I_m+[g(0,k_0)h(0,k_0)^{-1}]^*[g(0,k_0)h(0,k_0)^{-1}]\big]^{-1/2}, & k_0
\text{ odd},\\
\big[I_m+[h(0,k_0)^{-1}g(0,k_0)]^*[h(0,k_0)^{-1}g(0,k_0)]\big]^{-1/2}, & k_0
\text{ even},
\end{cases}
\end{align}
and hence,
\begin{align}
\alpha_{k_0} =
\begin{cases}
g(0,k_0)h(0,k_0)^{-1}\rho_{k_0}, & k_0 \text{ odd},\\
h(0,k_0)^{-1}g(0,k_0)\rho_{k_0}, & k_0 \text{ even}.
\end{cases}
\end{align}
Using \eqref{3.10} and \eqref{3.11}, one also obtains
$a_{k_0}=I_m+\alpha_{k_0}$ and $b_{k_0}=I_m-\alpha_{k_0}$.
Next, utilizing \eqref{4.14}, \eqref{4.16}, and \eqref{4.17}, one
computes,
\begin{align}
\begin{split}
g(z,k_0)h(z,k_0)^{-1} &= -[I_m+M_-(z,k_0)][a_{k_0}^* -
b_{k_0}^*M_-(z,k_0)]^{-1}\rho_{k_0}, \quad k_0 \text{ odd},
\\
h(z,k_0)^{-1}g(z,k_0) &= -\widetilde\rho_{k_0}[a_{k_0}^* -
b_{k_0}^*M_-(z,k_0)]^{-1}[I_m+M_-(z,k_0)], \quad k_0 \text{ even}.
\end{split}
\end{align}
Solving for $M_-(z,k_0)$, one then obtains
\begin{align}
M_-(z,k_0) =
\begin{cases}
2g(z,k_0)[b_{k_0}^*g(z,k_0) - \rho_{k_0}h(z,k_0)]^{-1}-I_m, & k_0
\text{ odd},\\
2[g(z,k_0)b_{k_0}^* - h(z,k_0)\widetilde\rho_{k_0}]^{-1}g(z,k_0)-I_m, &
k_0 \text{ even}. \label{5.51}
\end{cases}
\end{align}
The right-hand side of the above formula is well-defined for
sufficiently small $|z|$ since $b_{k_0}^*g(z,k_0) -
\rho_{k_0}h(z,k_0)$ for $k_0$ odd and $g(z,k_0)b_{k_0}^* -
h(z,k_0)\widetilde\rho_{k_0}$ for $k_0$ even are ${\mathbb{C}^{m\times m}}$-valued analytic
functions having invertible values at the origin,
\begin{align}
\begin{split}
b_{k_0}^*g(0,k_0) - \rho_{k_0}h(0,k_0) &=
(\alpha_{k_0}-I_m)\rho_{k_0}^{-1}h(0,k_0), \quad k_0 \text{ odd},
\\
g(0,k_0)b_{k_0}^* - h(0,k_0)\widetilde\rho_{k_0} &=
h(0,k_0)\widetilde\rho_{k_0}^{-1}(\alpha_{k_0}-I_m), \quad k_0 \text{ even}.
\end{split} \label{5.52}
\end{align}
Next, having $M_-(z,k_0)$ for sufficiently small $|z|$, one solves the equation
\begin{align}
h(z,k_0) = -\frac{1}{2z}
\begin{cases}
\rho_{k_0}^{-1}
[a_{k_0}^*-b_{k_0}^*M_-(z,k_0)][M_+(z,k_0)-M_-(z,k_0)]^{-1}
[I_m-M_+(z,k_0)],
& k_0 \text{ odd},\\
[I_m-M_+(z,k_0)][M_+(z,k_0)-M_-(z,k_0)]^{-1}
[a_{k_0}^*-M_-(z,k_0)b_{k_0}^*] \widetilde\rho_{k_0}^{-1}, & k_0 \text{
even},
\end{cases}
\end{align}
for $M_+(z,k_0)$ and obtains,
\begin{align}
M_+(z,k_0) &=
\begin{cases}
2[I_m+zg(z,k_0)]\big[I_m+z[b_{k_0}^*g(z,k_0)-\rho_{k_0} h(z,k_0)]\big]^{-1} -I_m, &
k \text{ odd},\\
2\big[I_m+z[g(z,k_0)b_{k_0}^*-
h(z,k_0)\widetilde\rho_{k_0}]\big]^{-1}[I_m+zg(z,k_0)] -I_m, & k_0 \text{ even}.
\label{5.53}
\end{cases}
\end{align}
The right-hand side of \eqref{5.53} is well-defined for
sufficiently small $|z|$ since both $I_m+z(b_{k_0}^*g(z,k_0)
-\rho_{k_0}h(z,k_0))$ and $I_m+z(g(z,k_0)b_{k_0}^*-
h(z,k_0)\widetilde\rho_{k_0})$ are ${\mathbb{C}^{m\times m}}$-valued analytic functions having
invertible values at the origin.
Finally, Theorem \ref{t5.1} (parts $(i)$, $(iv)$ and $(vi)$, $(ix)$)
implies that $M_\pm(z,k_0)$ for $z$ in some small neighborhood of
the origin uniquely determine Verblunsky coefficients
$\{\alpha_{k}\}_{k\in{\mathbb{Z}}}$.
{\it Case} $(ii)$. Suppose $k_0$ is odd. Then \eqref{3.133}, \eqref{4.14},
\eqref{4.15}, and
\begin{align}
2[I_m+zg(z,k_0)] &= [I_m+M_-(z,k_0)]W(z,k_0)^{-1}[I_m-M_+(z,k_0)]
\notag
\\ &\quad +
[M_+(z,k_0)-M_-(z,k_0)]W(z,k_0)^{-1} +
W(z,k_0)^{-1}[M_+(z,k_0)-M_-(z,k_0)] \notag
\\ & =
[I_m+M_+(z,k_0)]W(z,k_0)^{-1}[I_m-M_-(z,k_0)] \label{5.55}
\end{align}
imply the identity,
\begin{align}
& z\rho_{k_0} g(z,k_0-1)\rho_{k_0} \notag
\\
&\quad =
\frac12 \big[(I_m+\alpha_{k_0}^*)-(I_m-\alpha_{k_0}^*)M_+(z,k_0)\big]
W(z,k_0)^{-1} \big[(I_m+\alpha_{k_0})+M_-(z,k_0)(I_m-\alpha_{k_0})\big] \notag
\\
&\quad = \frac12 \big[[I_m-M_+(z,k_0)]+\alpha_{k_0}^*[I_m+M_+(z,k_0)]\big]
W(z,k_0)^{-1} \label{5.54}
\\
&\qquad \times \big[[I_m+M_-(z,k_0)]+[I_m-M_-(z,k_0)]\alpha_{k_0}\big]
\notag
\\
&\quad = \frac12 [-\Phi_+(z,k_0)+\alpha_{k_0}^*]
[I_m+M_+(z,k_0)]W(z,k_0)^{-1}[I_m-M_-(z,k_0)]
[-\Phi_-(z,k_0)^{-1}+\alpha_{k_0}] \notag
\\
&\quad = [\alpha_{k_0}^*-\Phi_+(z,k_0)] [I_m+zg(z,k_0)]
[\alpha_{k_0}-\Phi_-(z,k_0)^{-1}]. \notag
\end{align}
Moreover, \eqref{5.55} also implies
\begin{align}
&zg(z,k_0)[I_m+zg(z,k_0)]^{-1} = [I_m+zg(z,k_0)]^{-1}zg(z,k_0) =
I_m -[I_m+zg(z,k_0)]^{-1} \notag
\\ &\quad =
[I_m-M_-(z,k_0)]^{-1} \big[[I_m-M_-(z,k_0)][I_m+M_+(z,k_0)]-2W(z,k_0)\big]
[I_m+M_+(z,k_0)]^{-1} \notag
\\ &\quad =
[I_m-M_-(z,k_0)]^{-1}
\big[I_m+M_-(z,k_0)-M_+(z,k_0)-M_-(z,k_0)M_+(z,k_0)\big]
[I_m+M_+(z,k_0)]^{-1} \notag
\\ &\quad =
[I_m-M_-(z,k_0)]^{-1}[I_m+M_-(z,k_0)][I_m-M_+(z,k_0)][I_m+M_+(z,k_0)]^{-1}
\label{5.56}
\\ &\quad =
\Phi_-(z,k_0)^{-1}\Phi_+(z,k_0). \notag
\end{align}
Introducing the ${\mathbb{C}^{m\times m}}$-valued analytic functions $A(z,k_0)$ and
$B(z,k_0)$ by
\begin{align}
A(z,k_0)=I_m+zg(z,k_0) \,\text{ and }\, B(z,k_0)=z\rho_{k_0}
g(z,k_0-1)\rho_{k_0} - \alpha_{k_0}^*A(z,k_0)\alpha_{k_0}, \label{5.57}
\end{align}
one rewrites \eqref{5.54} as
\begin{align}
\Phi_+(z,k_0)A(z,k_0)\alpha_{k_0} + B(z,k_0) -
\Phi_+(z,k_0)A(z,k_0)\Phi_-(z,k_0)^{-1} +
\alpha_{k_0}^*A(z,k_0)\Phi_-(z,k_0)^{-1} = 0.
\end{align}
Multiplying both sides by $\Phi_+(z,k_0)$ on the right and utilizing \eqref{5.56}
then yields the Riccati-type equation for $\Phi_+(z,k_0)$,
\begin{align}
\Phi_+(z,k_0)A(z,k_0)\alpha_{k_0}\Phi_+(z,k_0) + B(z,k_0)\Phi_+(z,k_0)
- \Phi_+(z,k_0)zg(z,k_0) + \alpha_{k_0}^*zg(z,k_0) = 0. \label{5.59}
\end{align}
Since by \eqref{3.128} and \eqref{3.133} $\Phi_+(0,k_0)=0$ and by
\eqref{5.57}
\begin{align}
&zg(z,k_0) \underset{z\to 0}{\longrightarrow} 0, \quad
A(z,k_0) \underset{z\to 0}{\longrightarrow} I_m, \quad
B(z,k_0) \underset{z\to 0}{\longrightarrow} \alpha_{k_0}^*\alpha_{k_0}, \label{5.60}
\end{align}
Lemma \ref{l5.4} implies that equation \eqref{5.59} uniquely
determines the analytic function $\Phi_+(z,k_0)$ for $|z|$
sufficiently small.
Having $\Phi_+(z,k_0)$, one obtains $\Phi_-(z,k_0)^{-1}$ from \eqref{5.54}
for $|z|$ sufficiently small,
\begin{align}
\Phi_-(z,k_0)^{-1} = \alpha_{k_0} - [I_m+zg(z,k_0)]^{-1}
[\alpha_{k_0}^*-\Phi_+(z,k_0)]^{-1} z\rho_{k_0}g(z,k_0-1)\rho_{k_0}.
\label{5.61}
\end{align}
The right-hand side of \eqref{5.61} is well-defined since
$I_m+zg(z,k_0)$ and $\alpha_{k_0}^*-\Phi_+(z,k_0)$ are ${\mathbb{C}^{m\times m}}$-valued
analytic functions invertible at the origin.
Finally, Theorem \ref{t5.1} (parts $(i)$, $(v)$ and $(vi)$, $(x)$)
implies that $\Phi_\pm(z,k_0)^{\pm1}$ for $|z|$ sufficiently small
uniquely determine the Verblunsky coefficients $\{\alpha_{k}\}_{k\in{\mathbb{Z}}}$.
The case of $k_0$ even is proved similarly.
\end{proof}
In the subsequent result, $g^{(j)}$ and $h^{(j)}$ denote the corresponding quantities
\eqref{5.62b} and \eqref{5.62c} associated with the Verblunsky coefficients
$\alpha^{(j)}$, $j=1,2$.
\begin{theorem} \label{t4.6}
Assume Hypothesis \ref{h3.1} for two sequences $\alpha^{(1)}$,
$\alpha^{(2)}$ and let $k_0\in{\mathbb{Z}}$, $N\in{\mathbb{N}}$. Then for the full-lattice
problems associated with $\alpha^{(1)}$ and $\alpha^{(2)}$ the following
local uniqueness results hold:
\begin{enumerate}[$(i)$]
\item
If either $h^{(1)}(0,k_0)$ or $h^{(2)}(0,k_0)$ is invertible and
\begin{align}
\begin{split}
&\big\|g^{(1)}(z,k_0)-g^{(2)}(z,k_0)\big\|_{\mathbb{C}^{m\times m}}
+ \big\|h^{(1)}(z,k_0)-h^{(2)}(z,k_0)\big\|_{\mathbb{C}^{m\times m}} \underset{z\to 0}{=} o(z^N), \label{5.71} \\
& \, \text{then } \, \alpha^{(1)}_k = \alpha^{(2)}_k \,\text{ for }\,
k_0-N \leq k\leq k_0+N+1.
\end{split}
\end{align}
\item
If $\alpha^{(1)}_{k_0}=\alpha^{(2)}_{k_0}$, $\alpha^{(1)}_{k_0}$ is
invertible, and
\begin{align}
\begin{split}
&\big\|g^{(1)}(z,k_0-1)-g^{(2)}(z,k_0-1)\big\|_{\mathbb{C}^{m\times m}} +
\big\|g^{(1)}(z,k_0)-g^{(2)}(z,k_0)\big\|_{\mathbb{C}^{m\times m}} \underset{z\to 0}{=} o(z^N), \label{5.72} \\
& \, \text{then } \, \alpha^{(1)}_k = \alpha^{(2)}_k \,\text{ for }\,
k_0-N-1 \leq k\leq k_0+N+1.
\end{split}
\end{align}
\end{enumerate}
\end{theorem}
\begin{proof}
{\it Case} $(i)$. The result is implied by Theorem \ref{t5.2} (parts
$(i)$, $(iii)$ and $(v)$, $(vii)$) upon verifying that \eqref{5.51},
\eqref{5.53}, and \eqref{5.71} imply
\begin{align}
\begin{split}
\big\|M_+^{(1)}(z,k_0)-M_+^{(2)}(z,k_0)\big\|_{\mathbb{C}^{m\times m}} &\underset{z\to 0}{=} o(z^{N+1}),
\\
\big\|M_-^{(1)}(z,k_0)-M_-^{(2)}(z,k_0)\big\|_{\mathbb{C}^{m\times m}} &\underset{z\to 0}{=} o(z^N).
\end{split}
\end{align}
{\it Case} $(ii)$. The result is a consequence of Theorem \ref{t5.2} (parts
$(i)$, $(iv)$ and $(v)$, $(viii)$) upon verifying that Corollary
\ref{c5.4}, \eqref{5.57}, \eqref{5.59}, \eqref{5.61}, and
\eqref{5.72} imply
\begin{align}
\norm{\Phi_+^{(1)}(z,k_0)-\Phi_+^{(2)}(z,k_0)}_{\mathbb{C}^{m\times m}} +
\norm{\Phi_-^{(1)}(z,k_0)^{-1}-\Phi_-^{(2)}(z,k_0)^{-1}}_{\mathbb{C}^{m\times m}} \underset{z\to 0}{=} o(z^{N+1}).
\end{align}
\end{proof}
|
\section{Introduction}
\label{Sec:intro}
The fixed-node diffusion Monte Carlo (DMC) method is often the method of choice for
accurate computations of many-body systems~\cite{FMNR01}.
Since the scaling of DMC with the number of electrons $N$ is a modest $N^4$, the method
has been employed in recent years to accurately compute electronic properties of large
molecular and solid systems where conventional highly-correlated quantum chemistry approaches
are very difficult to apply. Unfortunately, for full-core atoms the computational cost of DMC increases
approximately~\cite{C86,HRL87} as $Z^{5.5-6.5}$ with the atomic number $Z$. Therefore, the
use of pseudopotentials is an essential ingredient in the application of DMC to complex
systems to reduce the effective value of $Z$ and significantly improve the efficiency
of the method.
The use of pseudopotentials in DMC poses however a problem since pseudopotentials are usually
non-local and the non-locality introduces a fermionic sign problem additional to the
one due to the anti-symmetry of the electronic wave function. The commonly adopted
solution is to ``localize'' the non-local pseudopotential on the trial wave function and use this
local potential in the DMC simulation~\cite{fahy,mitas}. Unfortunately,
the so-called locality approximation (LA) does not
ensure variationality and alternative schemes employing a different effective Hamiltonian were recently
introduced to overcome this difficulty~\cite{CFS05,C06}.
In the lattice regularized diffusion Monte Carlo (LRDMC) algorithm~\cite{CFS05}, both the Laplacian and the
non-local pseudopotentials are discretized such that the corresponding imaginary time
propagator $\langle \exp[-\tau {\cal H}] \rangle$ assumes non-zero
values only on a finite set of points, and the lattice Green function Monte Carlo algorithm can be employed,
that ensures variationality and stability all along the simulation~\cite{ceperley}.
Alternatively, another scheme which
is based on the standard DMC algorithm was developed~\cite{C06}. The latter exploits the discretization
of the propagator only in the part depending on the non-local pseudopotentials,
and a non-local effective Hamiltonian is defined in order to
fulfill the fixed node constraint.
Here, we show that this variational DMC scheme
is however not size consistent at finite time-steps. Indeed, the time-step error
strongly depends on the system size and, upon increasing the number of particles at fixed time-step,
the corresponding energies approach those given by DMC with LA.
In this paper, we explain how to cure this problem and
present a simple formulation of the algorithm
which is size-consistent and suffers at the same time from a smaller time-step error.
Moreover, we define a better discretization rule for the LRDMC effective Hamiltonian,
which reduces the lattice-space bias, remains size consistent as in the
original formulation,
and improves the efficiency of the method.
In Section~\ref{Sec:methods}, we briefly summarize the problems introduced by the use of
non-local pseudopotentials in the standard DMC and describe in detail the variational DMC algorithm
of Ref.~\onlinecite{C06} to treat pseudopotentials beyond the commonly used LA.
In Section~\ref{Size_consis}, we present our size-consistent variational approach to
non-local pseudopotentials in DMC and demonstrate its effectiveness on a series
of oxygen systems of increasing size.
In Section~\ref{Sec:lrdmc}, we briefly describe the LRDMC method, which is variational
by its own nature, and give a better prescription for the lattice regularization of the continuous Hamiltonian to
always guarantee a well defined and faster zero lattice-space extrapolation.
Finally, in Section~\ref{Sec:performance}, we discuss the behavior of the discretization error
of the different DMC algorithms (in time or space as appropriate)
and comment on the relative efficiency of the methods presented here.
\section{DMC and non-local pseudopotentials}
\label{Sec:methods}
In DMC, the projection to the ground state wave function of an Hamiltonian $\cal H$ is
performed by stochastically applying the operator $\exp[-\tau{\cal H}]$ to a trial wave
function $\Psi_{\rm T}$. If the projection is formulated in real space and importance
sampling introduced, the mixed distribution
$f({\bf R},t)=\Psi_{\rm T}({\bf R})\Psi({\bf R},t)$ is then propagated as
\begin{eqnarray}
f({\bf R}',t+\tau)=\int {\rm d} {\bf R}\, G({\bf R}',{\bf R},\tau)f({\bf R},t)\,,
\end{eqnarray}
where the importance sampling Green's function is defined as
\begin{eqnarray}
G({\bf R}',{\bf R},\tau)&=&\frac{\Psi_{\rm T}({\bf R}')}{\Psi_{\rm T}({\bf R})}\langle {\bf R}'|\exp[-\tau{\cal H}]|{\bf R}\rangle\,.
\end{eqnarray}
The fixed-node (FN) approximation is usually employed for fermionic systems to avoid the collapse
to the bosonic ground state.
In continuous systems, it is implemented by constraining the diffusion process
within the nodal pockets of the
trial wave function. For long times, the distribution $f({\bf R},t)$ approaches
$\Psi_{\rm T}({\bf R})\Psi_{\rm FN}({\bf R})$ where $\Psi_{\rm FN}({\bf R})$ is the ground state wave function
consistent with the boundary condition that it vanishes at the nodes of $\Psi_{\rm T}$. The FN energy is an
upper bound to the true fermionic ground state energy.
When a non-local potential ${\cal V}^{\rm NL}$ is employed to remove the core electrons, the off-diagonal elements
of the Hamiltonian in real space are generally non-zero and the standard DMC approach cannot be applied.
If we analyze the behavior of the propagator at short time-steps:
\begin{eqnarray}
\langle {\bf R}' | \exp[-\tau{\cal H}] | {\bf R} \rangle
\approx \delta_{{\bf R}',{\bf R}}-\tau\langle {\bf R}' | {\cal H} | {\bf R} \rangle\,,
\end{eqnarray}
we note that, while the diagonal elements can always be made positive by choosing $\tau$ small enough, the off-diagonal
elements are positive if and only if the off-diagonal elements of the Hamiltonian are non-positive. This condition is not
always met in the presence of non-local pseudopotentials, so the fermionic sign problem reappears even if one works
in the FN approximation.
Consequently, in addition to the FN approximation, the LA is commonly introduced where
the non-local potential ${\cal V}^{\rm NL}$ is replaced by a local quantity ${\cal V}^{\rm LA}$ obtained by ``localizing'' the potential
on the trial wave function:
\begin{eqnarray}
{\cal V}^{\rm LA}({\bf R})=\frac{\langle {\bf R} | {\cal V}^{\rm NL} | \Psi_{\rm T} \rangle}
{\langle {\bf R} |\Psi_{\rm T} \rangle}\,.
\end{eqnarray}
The DMC algorithm in the LA yields the FN ground state of the effective Hamiltonian
${\cal H}^{\rm LA}$ with the local potential ${\cal V}^{\rm LA}$ instead of the original non-local ${\cal V}^{\rm NL}$.
The fixed-node energy in the LA is equal to
\begin{eqnarray}
{\rm E}^{\rm LA}_{\rm FN}=\frac{\langle \Psi^{\rm LA}_{\rm FN} | {\cal H}^{\rm LA} | \Psi^{\rm LA}_{\rm FN} \rangle}{\langle \Psi^{\rm LA}_{\rm FN} | \Psi^{\rm LA}_{\rm FN} \rangle}\,,
\end{eqnarray}
and estimated by sampling the mixed distribution $\Psi_{\rm T}\Psi^{\rm LA}_{\rm FN}$ as
\begin{eqnarray}
{\rm E}^{\rm LA}_{\rm FN}
=\frac{\langle \Psi^{\rm LA}_{\rm FN} | {\cal H}^{\rm LA} | \Psi_{\rm T} \rangle}{\langle \Psi^{\rm LA}_{\rm FN} | \Psi_{\rm T} \rangle}
=\frac{\langle \Psi^{\rm LA}_{\rm FN} | {\cal H} | \Psi_{\rm T} \rangle}{\langle \Psi^{\rm LA}_{\rm FN} | \Psi_{\rm T} \rangle}\,.
\end{eqnarray}
Since $\Psi^{\rm LA}_{\rm FN}$ is the fixed-node ground state of ${\cal H}^{\rm LA}$ and not of the original
Hamiltonian ${\cal H}$, the mixed average energy of ${\cal H}$ is not equal to its expectation value on
the wave function $\Psi^{\rm LA}_{\rm FN}$,
\begin{eqnarray}
\frac{\langle \Psi^{\rm LA}_{\rm FN} | {\cal H} | \Psi^{\rm LA}_{\rm FN} \rangle}{\langle \Psi^{\rm LA}_{\rm FN} | \Psi^{\rm LA}_{\rm FN} \rangle}\,.
\end{eqnarray}
Therefore, $E^{\rm LA}_{\rm FN}$ is in general not an upper bound to the ground state of ${\cal H}$
and the variational principle may not apply.
\subsection{Beyond the locality approximation}
\label{Beyond_LA}
The lattice regularized diffusion Monte Carlo (LRDMC) algorithm was recently developed to overcome this
difficulty~\cite{CFS05} and then extended to the continuum formulation of DMC~\cite{C06}.
Both algorithms provide a variational scheme to treat non-local pseudopotentials in DMC by introducing an
effective Hamiltonian, different from the one used in the LA approximation, which provides an upper bound
to the ground state of the original Hamiltonian. We briefly describe here the algorithm in the framework
of continuum DMC.
We first apply a Trotter expansion for small time-steps to the importance sampling Green's function,
\begin{eqnarray}
G({\bf R}',{\bf R},\tau)
\approx\int {\rm d}{\bf R}''T^{\rm NL}({\bf R}',{\bf R}'',\tau) G^{\rm loc}({\bf R}'',{\bf R},\tau)\,,
\label{int_form}
\end{eqnarray}
where we have split the Hamiltonian into a local and a non-local operator. The propagator
$G^{\rm loc}({\bf R}',{\bf R},\tau)$ is equal to the drift-diffusion-branching Green's function for the local
component of the Hamiltonian,
\begin{eqnarray}
\frac{1}{(2\pi\tau)^{3N/2}}e^{-\left[{\bf R}'-{\bf R}-\tau {\bf V}({\bf R})\right]^2/2\tau}
e^{-\tau E_{\rm L}^{\rm loc}({\bf R}')}\,,
\label{G^loc}
\end{eqnarray}
where the velocity is defined as ${\bf V}({\bf R})=\nabla\Psi_{\rm T}({\bf R})/\Psi_{\rm T}({\bf R})$ and
$E_{\rm L}^{\rm loc}({\bf R})={\cal H}^{\rm loc}\Psi_{\rm T}({\bf R})/\Psi_{\rm T}({\bf R})$ is the local energy of the
local part of the Hamiltonian (kinetic ${\cal K}$ plus local potential ${\cal V}^{\rm loc}$).
The transition $T^{\rm NL}$ contains the non-local potential,
\begin{eqnarray}
T^{\rm NL}({\bf R}',{\bf R},\tau)&=&\frac{\Psi_{\rm T}({\bf R}')}{\Psi_{\rm T}({\bf R})}\langle {\bf R}'|\exp[-\tau{\cal V}^{\rm NL}]|{\bf R}\rangle\nonumber\\
&\approx& \delta_{{\bf R}',{\bf R}}-\tau V_{{\bf R}',{\bf R}}\,.
\label{T^NL}
\end{eqnarray}
where $V_{{\bf R}',{\bf R}}=\Psi_{\rm T}({\bf R}')/\Psi_{\rm T}({\bf R}) \langle {\bf R}'|{\cal V}^{\rm NL}|{\bf R}\rangle$.
In both the variational Monte Carlo and the standard DMC method with the LA approximation, one adopts a quadrature
rule with a discrete mesh of points, belonging to a regular polyhedron used
to evaluate the projection of the non-local component on a given trial wave function.
Consequently, the number of elements $V_{{\bf R}',{\bf R}}$ is finite
and the transition $T^{\rm NL}$ corresponds to the move of one electron on the grid obtained by considering the union of
the quadrature points generated for each electron and pseudoatom
(center of a non-local pseudopotential).
Moreover, in order to work with a small quadrature mesh, the
vertices of the polyhedron are defined in a frame rotated by $\theta$ and $\phi$,
the azimuthal and planar angle respectively, which are taken randomly
for each electron.
As discussed above, performing a transition based on $T^{\rm NL}$ poses however a problem since $T^{\rm NL}$ can be negative given that both
$\Psi_{\rm T}({\bf R}')/\Psi_{\rm T}({\bf R})$ and $\langle {\bf R}'|{\cal V}^{\rm NL}|{\bf R}\rangle$ can change sign.
A solution is to apply the FN approximation not only to $G^{\rm loc}$ but also to $T^{\rm NL}$ by keeping only the transition
elements which are positive:
\begin{eqnarray}
T^{\rm NL}_{\rm FN}({\bf R}',{\bf R},\tau)&=& \delta_{{\bf R}',{\bf R}}-\tau V^-_{{\bf R}',{\bf R}}\,.
\label{T^NL_FN}
\end{eqnarray}
where $V^\pm_{{\bf R}',{\bf R}}=[V_{{\bf R}',{\bf R}}\pm|V_{{\bf R}',{\bf R}}|]/2$. The discarded elements
are included in the so-called sign-flip term, which is then added to the diagonal local potential as
\begin{eqnarray}
{\cal V}_{\rm eff}^{\rm loc}({\bf R})={\cal V}^{\rm loc}({\bf R})+\sum_{{\bf R}'}V^+_{{\bf R}',{\bf R}}\,.
\label{V_eff^loc}
\end{eqnarray}
The resulting effective Hamiltonian ${\cal H}_{\rm eff}$ is therefore given by
\begin{eqnarray}
\langle{\bf R}|{\cal H}_{\rm eff}|{\bf R}\rangle&=& \langle{\bf R}|{\cal K}|{\bf R}\rangle+
{\cal V}_{\rm eff}^{\rm loc}({\bf R})\nonumber\\
\langle{\bf R}'|{\cal H}_{\rm eff}|{\bf R}\rangle&=& \langle{\bf R}'|{\cal V}^{\rm NL}|{\bf R}\rangle
\hspace*{1cm}{\rm if}\ V_{{\bf R}',{\bf R}}<0\,,
\label{H_eff}
\end{eqnarray}
and yields the same local energy as the original Hamiltonian ${\cal H}$. In contrast to
the LA Hamiltonian, its ground state energy is an upper bound to the ground state energy of the true
Hamiltonian. Therefore, the variational principle is recovered and, in addition, the use of ${\cal H}_{\rm eff}$
in combination with the $T_{\rm FN}^{\rm NL}$ transition cures the instabilities which are commonly observed in a DMC run with the
LA Hamiltonian and are due to the negative divergences of the localized potential on the nodes of $\Psi_{\rm T}$.
In the branching term of $G^{\rm loc}$ (Eq.~\ref{G^loc}), the local potential
${\cal V}^{\rm loc}$ is replaced by ${\cal V}_{\rm eff}^{\rm loc}$ (Eq.~\ref{V_eff^loc})
and the weights of the walkers are multiplied by an additional factor which enters in the normalization
of the transition $T_{\rm FN}$ and to order $\tau$ is equal to:
\begin{eqnarray}
\sum_{{\bf R}'}T_{\rm FN}^{\rm NL}({\bf R}',{\bf R}) \approx\exp\left(-\tau\sum_{{\bf R}'}V^-_{{\bf R}',{\bf R}}\right)\,.
\end{eqnarray}
The weights are therefore given by
\begin{eqnarray}
w=w_{\rm eff}^{\rm loc}\sum_{{\bf R}'}T_{\rm FN}^{\rm NL}({\bf R}',{\bf R}) =\exp\left[-\tau E_{\rm L}({\bf R})\right]\,,
\end{eqnarray}
where $E_{\rm L}({\bf R})={\cal H}_{\rm eff}\Psi_{\rm T}({\rm R})/\Psi_{\rm T}({\rm R})={\cal H}\Psi_{\rm T}({\rm R})/\Psi_{\rm T}({\rm R})$.
The basic algorithm as proposed in Ref.~\cite{C06} therefore consists of the following steps:
\begin{itemize}
\itemsep0ex
\item[1.] The walker drifts and diffuses
from ${\bf R}$ to ${\bf R}'$. The move is followed by an accept/reject step as in standard DMC.
\item[2.] The weight of the walker is multiplied by the branching factor
$\exp\left[-\tau\left(E_{\rm L}({\bf R}')-E_{\rm T}\right)\right]$ where the trial energy $E_{\rm T}$ has been
introduced.
\item[3.] The walker moves to ${\bf R}''$ according to the transition
probability $T^{\rm NL}_{\rm FN}({\bf R}'',{\bf R}')/\sum_{{\bf R}'''}T_{\rm FN}^{\rm NL}({\bf R}''',{\bf R}')$.
\end{itemize}
For large systems, the first step is implemented not by moving all the electron together but by sequentially drifting and diffusing
each electron and applying the accept/reject step after each single-electron (SE) move.
\section{Size-consistency}
\label{Size_consis}
In the move governed by the transition $T^{\rm NL}_{\rm FN}$, only one electron is displaced on the grid of the
quadrature points generated by considering all the pseudoatoms and all the electrons.
Therefore, for given time-step, the probability of a successful move will increase with the system size (i.e.\ the number of
electrons) and saturate to one for sufficiently
large systems. In this limit, the effect of the move will become independent of the system size and lead to one
electron being displaced at each step. Therefore, for sufficiently large systems, the overall
impact of the non-local move will decrease and the algorithm will effectively behaves more and more like in the LA procedure.
To demonstrate the size-consistency problem of the algorithm as originally formulated in Ref.~\cite{C06}, we consider a series of systems
consisting of an increasing number $M$ of oxygen atoms aligned 30~$\AA$ apart.
The oxygen atom is described by an $s$-non-local energy-consistent Hartree-Fock pseudopotential~\cite{BFD07}.
The trial wave function is of the Jastrow-Slater type with a single determinant expressed on a cc-pVDZ
basis~\cite{BFD07} and a Jastrow factor which includes electron-electron and
electron-nucleus terms~\cite{FU96}. All Jastrow and orbital parameters are optimized in energy
minimization~\cite{UTFSH07} for a single atom and the wave function of a system with more than one oxygen is
obtained by replicating the wave function of one atom on the other centers. In Fig.~\ref{fig:jwalkalize},
we plot the acceptance of the $T_{\rm FN}^{\rm NL}$ move as a function of time-step for systems containing
1, 2, 4, 8, 16, and 32 oxygen atoms. For each system size, the probability goes to zero at small time-steps
and increases for larger values of $\tau$ as expected from the expression of $T_{\rm FN}^{\rm NL}$ (Eq.~\ref{T^NL_FN}).
The acceptance increases with the size of the system; as a function of the time-step, it approaches its asymptotic value of one more quickly for the larger systems.
\begin{figure}[t]
\centering
\includegraphics[width=1.0\columnwidth]{jwalkalize.pdf}
\caption{Acceptance of the $T_{\rm FN}^{\rm NL}$ move as a function of time-step. The increasing dotted curves
correspond to systems with 1, 2, 4, 8, 16 and 32 oxygen atoms aligned at a distance of 30~$\AA$ from each other.
The dotted curves are obtained with the algorithm of Ref.~\onlinecite{C06} while the lowest continuum curve is obtained
with the size-consistent DMC algorithm (version 1) we propose. We only show the size-consistent curve with 1 atom as it is
indistinguishable from the ones obtained for the larger systems.}
\label{fig:jwalkalize}
\end{figure}
To better understand the overall behavior of the algorithm with increasing system size, we also analyze
the FN energy as a function of time-step.
We are interested in comparing the results obtained with the conventional LA approach and with the algorithm of
Ref.~\cite{C06} described in the previous section. For a more meaningful and clear comparison with conventional
DMC with the LA which employs a symmetrized branching factor, we modify the original algorithm as described in
the previous section to also use a symmetrized branching factor:
\begin{eqnarray}
\exp\left[-\tau\left(E_{\rm L}({\bf R})+E_{\rm L}({\bf R}')\right)/2\right]
\label{sym_branching}
\end{eqnarray}
where ${\bf R}$ and ${\bf R}'$ are the coordinates before the drift-diffusion move of the first electron and after
the drift-diffusion move of the last electron, respectively, if the electrons are displaced subsequently.
Such a simple modification is allowed as it only entails a different time-step error, which
we actually find to be significantly smaller than the one obtained with the algorithm of Ref.~\cite{C06} as
we detail in the Section~\ref{Sec:performance}.
\begin{figure}[t]
\centering
\includegraphics[width=1.0\columnwidth]{eofN.pdf}\\
\includegraphics[width=1.0\columnwidth]{eoftau.pdf}
\caption[]{Upper panel: DMC FN energy per atom for systems of $M$ isolated
oxygen atoms for $\tau=0.1$. For given time-step, the results of the algorithm
of Ref.~\cite{C06} (red circles) approach those of the LA (blue triangles)
upon increasing the system size, whereas the present algorithm (DMC version 1,
green squares) gives values independent of $M$. In the three algorithms, the
branching factor is updated after having moved all the electrons (AE
branching).
Lower panel: time-step dependence of the DMC FN energy for the same three
algorithms for $M=1$ (filled symbols) and $M=32$ (open symbols). The algorithm
of Ref.~\cite{C06} has a problematic extrapolation to zero time-step for large
enough system size. For $M=32$, the linear extrapolation (red curve with open
symbols) is consistent, as expected, with the corresponding result for $M=1$
(red curve with filled circles). However a better $\chi^2$ would be obtained
with a quadratic extrapolation, which in turn would require a sudden upturn
at very small $\tau$ to recover the correct zero time-step limit.
}
\label{fig:sizeconsistency_e}
\end{figure}
From the results reported in Fig.~\ref{fig:sizeconsistency_e}, we observe that
the FN energies obtained with the algorithm of Ref.~\cite{C06} significantly
increase with $M$ and approach the energies obtained with the LA algorithm.
Already with a system with 32 oxygen atoms, the FN energies at
$\tau=0.1$ obtained with these two approaches become equivalent within
the error bars. The energies given by the two algorithms must however
extrapolate to different values as the time-step goes to zero\cite{C06}.
The lower panel of Fig.~\ref{fig:sizeconsistency_e} shows that for $M=32$,
in particular, they have to depart from each other within a tiny time-step
interval near the origin.
Because of this behavior, the algorithm of Ref.~\cite{C06} is bound to have a
problematic extrapolation to the zero time-step limit for large enough systems.
\subsection{Size-consistent formulations: version 1}
To address this problem, the original algorithm of Ref.~\cite{C06} can be easily reformulated in a size-consistent manner by observing that the
transition $T^{\rm NL}$ (Eq.~\ref{T^NL}) can be factorized as
\begin{eqnarray}
T^{\rm NL}({\bf R}',{\bf R},\tau)&=&\frac{\Psi_{\rm T}({\bf R}')}{\Psi_{\rm T}({\bf R})}\prod_{i=1}^N\langle {\bf r}_i'|e^{-\tau{\nu}^{\rm NL}}|{\bf r}_i\rangle\nonumber\\
&=&\prod_{i=1}^N\frac{\Psi_{\rm T}({\bf r}'_1\ldots{\bf r}'_i,{\bf r}_{i+1}\ldots)}{\Psi_{\rm T}({\bf r}'_1\ldots{\bf r}'_{i-1},{\bf r}_i\ldots)}\langle {\bf r}_i'|e^{-\tau{\nu}^{\rm NL}}|{\bf r}_i\rangle\nonumber\\
&=&\prod_{i=1}^N\left(\delta_{{\bf r}'_i{\bf r}_i}-\tau v_{{\bf r}'_i{\bf r}_i}\right)\nonumber\\
&=&\prod_{i=1}^N t^{\rm NL}({\bf r}'_i,{\bf r}_i,\tau)\,,
\end{eqnarray}
where $\nu^{\rm NL}$ is the non-local potential acting only on one electron due to all the atomic centers so that
the total non-local potential is given by the sum over the electrons,
$\langle {\bf R}'|{\cal V}^{\rm NL}|{\bf R}\rangle=\sum_i \langle {\bf r}_i'|{\nu}^{\rm NL}|{\bf r}_i\rangle$.
We defined the matrix element $v_{{\bf r}'_i{\bf r}_i}$ as
\begin{eqnarray}
v_{{\bf r}'_i{\bf r}_i}=\frac{\Psi_{\rm T}({\bf r}'_1\ldots{\bf r}'_i,{\bf r}_{i+1}\ldots)}{\Psi_{\rm T}({\bf r}'_1\ldots{\bf r}'_{i-1},{\bf r}_i\ldots)}\langle {\bf r}_i'|{\nu}^{\rm NL}|{\bf r}_i\rangle\,.
\end{eqnarray}
The transition $t^{\rm NL}({\bf r}'_i,{\bf r}_i,\tau)$ displaces the $i$-th electron over the grid of quadrature points
generated by considering only the $i$-th electron and all pseudoatoms which host the $i$-th electron in their core region.
The FN approximation is applied separately to each single-electron
transition as
\begin{eqnarray}
t_{\rm FN}^{\rm NL}({\bf r}',{\bf r},\tau)=\delta_{{\bf r}'{\bf r}}-\tau v^-_{{\bf r}'{\bf r}}\,,
\end{eqnarray}
where $v^\pm_{{\bf r}'{\bf r}}$ is defined in analogy to the case of the total non-local potential so that
only the positive transition elements are kept in the transition matrix.
In this formulation, the
third step of the DMC algorithm detailed above consists of a loop over the electrons where each
electron is subsequently moved according to the single-electron transition, $t_{\rm FN}^{\rm NL}$.
Therefore, while in the original algorithm of Ref.~\cite{C06} the
configuration generated in the $T_{\rm FN}^{\rm NL}$ step differs from the starting configuration
only in the coordinate of one electron, the configuration resulting from this size-consistent move will
generally change in more than one electronic coordinate and the number of electrons being moved will increase
with the size of the system.
To understand that the drift-diffusion-branching steps in the original algorithm do not need to be modified, we observe that
the expression of the effective Hamiltonian ${\cal H}_{\rm eff}$ we are working with is the same as in
Eq.~\ref{H_eff} in the limit of $\tau$ going to zero.
In particular, the sign-flip term obtained by summing all discarded terms $v^+_{{\bf r}'{\bf r}}$
over all the electrons is equal to the sign-flip term in ${\cal V}_{\rm eff}^{\rm loc}$ (Eq.~\ref{V_eff^loc})
to zero-order in $\tau$. Similarly, we have that, to order $\tau$,
\begin{eqnarray}
\label{v1}
\prod_{i=1}^N\sum_{{\bf r}'_i}t_{\rm FN}^{\rm NL}({\bf r}'_i,{\bf r}_i,\tau)
\approx \sum_{{\bf R}'}T_{\rm FN}^{\rm NL}({\bf R}',{\bf R},\tau)\,,
\end{eqnarray}
and we recover the same branching factor as in the original algorithm. Therefore, both algorithms
extrapolate to the same limit at zero time-steps. We will refer to this improved algorithm as ``DMC version 1'', to
distinguish it from another size-consistent version we will define later in this section.
We stress here that by ``version 1'' we do not only mean the use of the product of
single-particle $t_{\rm FN}^{\rm NL}$ in step 3, but also the symmetrization of the weights in step 2,
as described in Eq.~\ref{sym_branching}, where the initial configuration is taken before the diffusion process (step 1),
and the final is the one after step 1.
The acceptance as a function of time-step using the size-consistent DMC
algorithm (version 1) is shown in Fig.~\ref{fig:jwalkalize}. We only report
the result obtained with $M=1$ as the curves for the other system sizes
are exactly equivalent within statistical error. This finding can
be easily understood since the probability of moving a given electron
on the grid generated by considering all centers will be practically
the same as the probability computed using only the atom close to the
electron as all other centers are at least 30 $\AA$ far apart. Therefore,
in the new size-consistent algorithm, when more atoms are added to increase
the size of the system, the loop over the electrons will ensure that each
electron attempts a move around its closest center. The acceptance remains
therefore constant as more oxygen atoms are added. In a more realistic
systems (e.g.\ with closer oxygen atoms), there will be a weak dependence
on the size of the system but, after enough atoms have been added for most
electrons to experience an equivalent environment, the acceptance will
become independent on the system size given the short-range nature of
the non-local components of the pseudopotentials.
The FN energies obtained for the oxygen systems with this size-consistent
algorithm are compared in Fig.~\ref{fig:sizeconsistency_e} with the
results of the LA and of the original algorithm.
We observe that, as expected, the FN energies of the size-consistent
algorithm extrapolate to the same value as the original algorithm as
$\tau$ goes to zero. On the other hand, while the size-consistent FN energies
are close to the values obtained with the original method for the smallest,
one-atom system, the FN results obtained by the two methods depart from
each other as the system size increases. Importantly, the FN energies of
the size-consistent scheme do not approach the LA results for large systems
at finite $\tau$ and their extrapolation to zero time-step is therefore as
smooth for large as for small systems.
\subsection{Size-consistent formulations: version 2}
An alternative scheme to address the size-consistent problem of the original algorithm of Ref.~\cite{C06}
can be obtained through a different route by starting from Eq.~\ref{T^NL},
and breaking it up in $N$ terms with time-step of $\tau/N$, such that:
\begin{equation}
\label{v2}
\sum_{{\bf R}'} T_{\rm FN}^{\rm NL}({\bf R}',{\bf R},\tau) = \sum_{{\bf R}_1 \cdots {\bf R}_{N}}
\prod_{i=1}^N T_{\rm FN}^{\rm NL}({\bf R}_i,{\bf R}_{i-1},\tau/N)\,,
\end{equation}
with ${\bf R}_N={\bf R}'$, ${\bf R}_0={\bf R}$, and the sum over the quadrature points sampled by the
chain $\{ {\bf R}_0,\ldots,{\bf R}_i,\ldots,{\bf R}_N \}$
generated during the random walk. This is another way to evaluate the quantity in Eq.~\ref{v1}.
The difference is that
Eq.~\ref{v2} involves a product of $N$ all-electron factors, while Eq.~\ref{v1} is a
factorization of $N$ single-electron terms. Both will avoid the saturation of the
acceptance probability of the non-local Green's function $T_{\rm FN}^{\rm NL}$,
and therefore they will ensure a size-consistent time-step error.
Since Eq.~\ref{v2} requires the calculation of \emph{all} matrix elements
$V_{{\bf R}',{\bf R}}$ each time, it is more convenient to split the $N$ factors in such a way that
the diffusion move involving the $i$-th electron could be placed between the $i-1$-th and $i$-th factor,
and the corresponding branching weight
updated as a product of subsequent single-electron components:
\begin{eqnarray}
\prod_{i=1}^N\,\exp\left[-\frac{\tau}{N} E_{\rm L}({\bf r}'_1\ldots{\bf r}'_i,{\bf r}_{i+1}\ldots,{\bf r}_N)\right]\,,
\end{eqnarray}
where we can exploit the knowledge of $V_{{\bf R}',{\bf R}}$ to compute also $E_{\rm L}$ for every single-electron move.
We will call this algorithm ``DMC version 2''~\cite{rqmc}. It consists of the following steps:
\begin{itemize}
\itemsep0ex
\item[1.] Diffusion move of the $i-$th electron.
\item[2.] The weight of the walker is multiplied by the branching factor
$\exp\left[-\frac{\tau}{N} E_{\rm L}({\bf r}'_1\ldots{\bf r}'_i,{\bf r}_{i+1}\ldots,{\bf r}_N)\right]$.
\item[3.] The walker moves to ${\bf R}''$ according to the transition
probability $T^{\rm NL}_{\rm FN}({\bf R}'',{\bf R}',\tau/N)/\sum_{{\bf R}'''}T_{\rm FN}^{\rm NL}({\bf R}''',{\bf R}',\tau/N)$
which involves \emph{all} the electrons.
\end{itemize}
In contrast to the original algorithm, and the ``DMC version 1'',
these three steps need to be performed inside a loop over the electrons.
In the ``version 2'' formulation of the DMC size-consistent algorithm,
each electron drifts and diffuses in a time $\tau$ and the branching factor is
updated at each SE move with the total local energy $E_{\rm L}$ and
time $\tau/N$ in the exponent. After each SE branching update,
a non-local transition is performed with
$T^{\rm NL}_{\rm FN}({\bf R}',{\bf R},\tau/N)$,
where one electron among {\it all} electrons is displaced over the grid
of quadrature points obtained by considering all electrons and all pseudoatoms.
Therefore, the electron displaced in the non-local move may differ from
the electron which is currently being moved in the drift-diffusion step.
\section{LRDMC and non-local pseudopotentials}
\label{Sec:lrdmc}
The main difference between the effective DMC Hamiltonian reported in Eq.~\ref{H_eff}
and the LRDMC one is the kinetic operator ${\cal K}$. In the LRDMC approach, ${\cal K}$ is
replaced by a discretized Laplacian and treated on the same footing as ${\cal V}^{\rm NL}$.
In the original formulation,
the discretized Laplacian is a linear combination of two discrete operators
with incommensurate lattice spaces $a$ and $a'$,
introduced to sample densely the continuous space by performing discrete moves
whose length is either $a$ or $a'$.
This method can be simplified by noticing that all the continuous
space can be visited using only a single displacement length $a$,
provided we randomize the direction of the Cartesian coordinates each time
the electron positions are updated.
The randomization of the lattice mesh is similar to the well established approach used
to perform the angular integration in the non-local part of the pseudopotential~\cite{fahy}.
Therefore, in the LRDMC approach, we can extend the definition of
the kinetic part by including both the discretized Laplacian and the
non-local part of the pseudopotentials. The total non-local operator reads:
\begin{equation}
\label{k_operator}
{\cal K}^a=
-\sum\limits_{i=1}^N \Delta^a_i(\theta_i,\phi_i)/2+ {\cal V}^{NL} \,,
\end{equation}
where $\Delta^a_i(\theta_i,\phi_i)$ is the Laplacian acting on the
$i$-th electron and discretized to second order so that $\Delta^a_i (\theta_i,\phi_i) = \Delta_i + O(a^2)$.
The discretized Laplacian is computed in a frame rotated by the angles
$\theta_i$ and $\phi_i$, which are chosen randomly and independently of
the ones used to compute ${\cal V}^{NL}$.
In this formulation, we need to evaluate only 6 off-diagonal elements of the Green's function instead of
12 as in the original algorithm, gaining a speed-up of a factor of 2 in full-core calculations and of
${(12+\mathrm{n_\mathrm{quad}})}/{(6+\mathrm{n_\mathrm{quad}})}$ with pseudopotentials, where
$\mathrm{n_\mathrm{quad}}$ is the number of quadrature points per electron~\cite{heavy}.
We notice that, with this simplification, the LRDMC error
in the extrapolation to the continuous limit depends on a single
parameter $a$, and the method can therefore be compared fairly with the DMC
approach where the discretization of the diffusion process also
depends on a single scale, i.e. the
time-step $\tau$.
In the LRDMC choice of the Hamiltonian, we further regularize the
single-particle operator ${\nu}$,
defined as the electron-ion Coulomb interaction in full-core atoms or the local part ${\nu}^{\rm loc}$ of the pseudopotential,
so that
${\nu}({\bf r}_i) \to {\cal V}_i^a ({\bf R})$ as
\begin{equation}
{\cal V}_i^a({\bf R}) = {\nu} ({\bf r}_i) - { \frac{( \Delta_i - \Delta_i^a)
\Psi_T({\bf R})}{2 \Psi_T({\bf R})} }\,,
\label{defvc}
\end{equation}
when acting on the $i$-th electron. The single-particle operator ${\nu}$
acquires therefore a many-body term and ${\cal V}_i^a({\bf R})$
depends on the all-electron configuration ${\bf R}$. The total potential term is then given by
\begin{equation}
{\cal V}^a = \sum\limits_{i=1}^N{\cal V}_i^a + {\cal V}_{ee} + {\cal V}_{nn}\,,
\label{pot_tot}
\end{equation}
where no regularization is employed in the electron-electron ${\cal V}_{ee}$
and ion-ion ${\cal V}_{nn}$ Coulomb terms.
This lattice regularization leads to an approximate
Hamiltonian ${\cal H}^a = {\cal K}^a + {\cal V}^a$
which converges to the exact Hamiltonian
as ${\cal H}^a = {\cal H}+a^2 \Delta {\cal H}$
for $a\to 0$, where we denote with $a^2 \Delta {\cal H}$ the
$O(a^2)$ LRDMC error on ${\cal H}$.
The lattice Green's function Monte Carlo algorithm can then be employed
to sample exactly the lattice regularized Green's function, $\Lambda - {\cal H}^a$, and project
the trial wave function $\Psi_T$ to the approximate ground state $\Psi_a^\mathrm{LRDMC}$
which fulfills the fixed-node constraint based on $\Psi_T$,
in complete analogy to the DMC framework~\cite{CFS05}.
Note that, since the
spectrum of ${\cal H}^a$ is not bounded from above, we need to take
the limit $\Lambda \rightarrow \infty$,
which can be handled with no loss of
efficiency as described in Ref.~\onlinecite{lambda}.
The usual DMC Trotter breakup results in a time-step error,
while the LRDMC formulation yields a lattice-space error, but both approaches
share the same upper bound property and converge to the same projected FN energy
in the limit of zero time-step and
lattice-space, respectively.
Since the discretized Laplacian and the non-local potential are treated on the same footing, and the
sampling of the Green's function is based on a sequence of single-particle moves generated both
from the Laplacian and the non-local part, the LRDMC is intrinsically size-consistent (in the sense
previously discussed for the DMC algorithm), and no modification is necessary to make the lattice-space bias independent
of the system size. It will depend however on the quality of the trial wave function in the way detailed below.
\subsection{Small $a^2$ correction for good trial function }
The regularization of the potential (Eq.~\ref{defvc})
in the definition of the lattice Hamiltonian ${\cal H}^a$
implies that the correction $\Delta {\cal H}$ satisfies:
\begin{equation} \label{p1}
\Delta {\cal H} | \Psi_T \rangle =0.
\end{equation}
Using this property, we can estimate the leading-order error of the lattice regularization
by simple perturbation theory as
\begin{eqnarray}
E^a&=& E^0 + a^2 \langle \Psi_0^\mathrm{LRDMC} | \Delta {\cal H} | \Psi_0^\mathrm{LRDMC} \rangle
\nonumber \\
&=& E^0 + a^2 \langle \Psi_0^\mathrm{LRDMC} - \Psi_T | \Delta {\cal H} | \Psi_0^\mathrm{LRDMC}
-\Psi_T \rangle
\nonumber \\
&=& E^0 + O( a^2 | \Psi_0^\mathrm{LRDMC} - \Psi_T |^2 ) \label{smalla}\,,
\end{eqnarray}
where $E^a$ is the expectation value of the Hamiltonian ${\cal H}^a$ on the
approximate FN ground state $\Psi_a^\mathrm{LRDMC} $ and $E^0$ the estrapolated
value as $a\rightarrow 0$.
Thus, the approach to
the continuous limit is particularly fast for good trial functions,
namely for $\Psi_T$ close to the ground state solution, since
$\Psi_0^\mathrm{LRDMC}$ is a state with lower energy than $\Psi_T$ and
has to approach the ground state at least as $\Psi_T$ does.
The leading corrections to the continuous limit
are quadratic in the wave function error. This property is not easily generalized
to the usual DMC method and, to our knowledge, has not been
established so far.
\subsection{Well defined lattice regularization}
As in any lattice model, the Hamiltonian ${\cal H}^a$
has a finite ground state energy
only if the potential ${\cal V}^a$ is always limited from below.
If ${\cal V}^a( {\bf R}_0) = -\infty $ for some configuration ${\bf R}_0$, the
variational state $\Psi({\bf R}) = \delta_{{\bf R},{\bf R}_0}$ will have unbounded negative
energy expectation value and the ground state
energy of ${\cal H}^a$ is not defined.
Unfortunately, the regularized potential ${\cal V}^a({\bf R})$ in Eq.~(\ref{defvc}) is not bounded from below when ${\bf R}$ belongs to
the $(3N-1)$-dimensional nodal surface ${\cal N}$ defined by the equation $ \Psi_T({\bf R})=0$.
To cure these divergences, we need to be able to establish when a configuration is close
to the nodal surface.
In the lattice regularized formulation, we can assign an electron position
${\bf r}_i$ to the nodal surface, i.e.\ $ {\bf r}_i \in {\cal N }_a$,
if $ \Psi_T ( {\bf r}_i + a \vec \mu)$ has the opposite sign of
$ \Psi_T ({\bf r}_i)$ for at least one of the six points used to evaluate the finite difference Laplacian
(i.e. $\vec \mu$ is one of the six unit vectors $\pm \hat x, \pm \hat y$
or $\pm \hat z$ of the reference frame randomly oriented according to the
angles $\theta_i$ and $\phi_i$).
${\cal N}_a$ correctly defines the nodal surface ${\cal N}$ in the limit $a\to 0$.
With this definition of nodal surface, we can
modify ${\cal V}_i^a$ so that it remains
finite when $ {\bf r}_i \in {\cal N}_a$:
\begin{equation} \label{vpota}
\tilde {\cal V}_i^a ( {\bf R} ) = \left\{
\begin{array}{cc}
\mathrm{Max} \left[ {\nu} ( {\bf r}_i) , { \cal V}_i^a ( {\bf R}) \right] &
{\rm if } ~ {\bf r}_i \in { \cal N }_a \\
{\cal V}_i^a( {\bf R}) & {\rm otherwise }
\end{array}\,.
\right.
\end{equation}
If $ {\bf r}_i \notin {\cal N}_a$,
we use the original LRDMC definition of ${\cal V}_i^a$ since
${\cal V}_i^a$ remains finite even when an electron approaches a nucleus
for trial functions which satisfy the electron-ion
cusp conditions. If $ {\bf r}_i \in {\cal N}_a$, we need to distinguish two cases. If the electron
is not close to a nucleus, the regularized ${ \cal V}_i^a$ can diverge negatively
while ${\nu}({\bf r}_i)$ remains finite and, according to Eq.~\ref{vpota},
the potential $\tilde {\cal V}_i^a$ coincides with $\nu$.
If the electron is close to a nucleus in a full-core calculation,
both ${ \cal V}_i^a$ and ${\nu}({\bf r}_i)$ diverge, so we need to further
regularize ${\nu} ({\bf r}_i)$ in the right hand side of Eq.~(\ref{vpota})
and use an expression bounded from below. In this particular case, we choose to
replace the divergent electron-ion contribution $ -Z/ | {\bf r}_{in} |$
in ${\nu} ({\bf r}_i)$
with $ - Z/a $ whenever the electron-ion distance $ |{\bf r}_{in}| < a $.
If we employ the regularized potential $\tilde {\cal V}^a$ in the Hamiltonian ${\cal H}^a$, we no longer satisfy Eq.~({\ref{p1})
and, in principle, it is not possible to compute $E^a$ by averaging the local energy ${\cal H}\Psi_{\rm T}/\Psi_{\rm T}$.
However, the use of $\tilde {\cal V}^a$ introduces only
negligible errors in the computation of $E^a$ because the regularization is adopted only in
a region of volume $S \times a$,
where $S$ is the area of the nodal surface ${\cal N }$. Since
both the trial and the LRDMC wave function vanish $\simeq a$
close to the nodal surface,
the finite lattice error corresponds to averaging
$ ({\cal H}^a-{\cal H}) \Psi_T/\Psi_T (\propto a$)
over $\Psi_T \Psi_{FN} (\propto a^2$)
in a nodal region of extension $\propto a$.
If we collect these contributions,
we find that the present regularization
introduces a bias in the nodal region which vanishes as $ a^4$
for $a \to 0$ and is always negligible compared to the dominant
contribution $ O( a^2 | \Psi_0 - \Psi_T |^2 )$.
Moreover, since the regularization in Eq.~(\ref{vpota}) acts independently
on each electron,
it does not affect the size-consistent character of the algorithm, and
the energy of $N$ independent atoms at large distances is equal to
$N$ times the energy of a single atom.
Therefore, we did not perform any LRDMC calculations for the oxygen systems since
the energy per atom as a function of $a$ is exactly independent of $N$.
\section{Performance of the proposed methods}
\label{Sec:performance}
An important point to address is the efficiency of our revised techniques.
This involves not only the computational cost per Monte Carlo step, but
also the elimination of the discretization error (in time or space, as
appropriate) by extrapolation to the continuum limit. Indeed, a smaller
and smoother bias enhances the overall efficiency, as does the knowledge
of the leading term in the discretization parameter.
\subsection{Time-step error}
We study the time-step error on the FN energy computed with the various algorithms discussed above,
using the Oxirane molecule (C$_2$H$_4$O) as a test case.
Our aim here is in particular to assess the reduction of the time-step error
with respect to the original algorithm~\cite{C06}.
In the DMC ``version 1'' this reduction is due to the symmetrization of the weights,
while in the DMC ``version 2'' it is due to the
update of the branching factor after single-electron moves.
We employ non-local energy-consistent Hartree-Fock
pseudopotentials~\cite{BFD07} for the oxygen and the carbon atoms in combination with the corresponding cc-pVDZ basis
sets, and construct two single-determinant Jastrow-Slater wave functions of different quality. The first wave function
is built from B3LYP orbitals and a very simple electron-electron Jastrow factor of the form $b [1-\exp(-\kappa r_{ij})]/\kappa$,
where $b=1/2$ or $1/4$ for antiparallel- and parallel-spin electrons, respectively. The parameter $\kappa$ is optimized in
energy minimization and is equal to 1.91.
The second wave function is characterized by a more sophisticated Jastrow factor comprising of
electron-electron, and electron-nucleus terms, and all orbital and Jastrow parameters in the
wave function are optimized in energy minimization.
The top panel of Fig.~\ref{fig:Oxirane} shows results
obtained with the simple wave function. Consistently with previous
studies on the water molecule~\cite{needs-water}, the LA energies
extrapolate to a lower value (not necessarily variational)
than the original algorithm of Ref.~\cite{C06}, with a
smaller time-step error; symmetrization of the branching factor
in the original algorithm is already sufficient to reduce
the time-step error down to a value similar to that found in the LA.
As expected, given the small size of the system considered, the
original and the size-consistent algorithm give nearly identical
results, as shown here for its version 1 with AE branching.
The main result shown in the top panel of Fig.~\ref{fig:Oxirane} is the remarkable reduction
of the time-step error obtained with a SE branching factor. The data shown
in the Figure refer to version 2 of the size-consistent algorithm.
We also mention, without reporting the data, that
when the branching factor is updated after SE moves, the symmetrization of
the local energy in the exponent does not improve the time-step error
significantly.
\begin{figure}[t]
\includegraphics[width=1.0\columnwidth]{oxirane_simplewf.pdf}\\
\includegraphics[width=1.0\columnwidth]{oxirane_goodwf.pdf}
\caption[]{FN energies as a function of time-step for the Oxirane (C$_2$H$_4$O) molecule, obtained using a simple (top) and a
more sophisticated (bottom) trial wave function. We employ different schemes, i.e. the original algorithm
as in Ref.~\cite{C06} and with an improved symmetrized branching factor (``sym''), the two size-consistent
approaches we proposed (``DMC version 1'', and ``DMC version 2''), the LA approach, and the LRDMC method.
The lattice-space has been mapped into the time-step via the relation $\tau = 0.6 ~ a^2$, which
guarantees the same autocorrelation time between the ``DMC version 1'' and the ``LRDMC'' method
in this particular case.}
\label{fig:Oxirane}
\end{figure}
The improvement obtained with a SE branching factor, however, is strongly
dependent on the quality of the trial function. The lower panel of
Fig.~\ref{fig:Oxirane} shows results obtained with the more sophisticated
wave function. We still find a lower energy with the LA (with its possibly
problematic behavior at very small time-step), and a large time-step error
with the original algorithm of Ref.~\cite{C06}. All the other cases, however,
display similar behavior, or at least comparable quality, in terms of the
time-step error.
Also the LRDMC energy values are reported in Fig.~\ref{fig:Oxirane}, where the lattice-space
has been converted into time-step based on the equal auto-correlation time between
Monte Carlo generations in the DMC and LRDMC algorithms. This is the fairest mapping
since it keeps the final statistical error equivalent for the same sample length.
In this case, it gives $\tau \sim 0.6 ~ a^2$.
One can see that the LRDMC
energies are always converging from below in a monotonic way, usually
easier to extrapolate than the corresponding DMC energies.
In order to make a more quantitative analysis of the predictions reported
in Section~\ref{Sec:lrdmc} for the lattice-space error, we studied
the lattice-space extrapolation of the Oxirane molecule with the DFT-B3LYP Slater determinant,
and Jastrow factors going from the simple
2-body one, to the most complicated comprising of one-, two-, and three-body terms.
The results are reported in Fig.~\ref{fig:lrdmc}. For good trial wave functions,
a reliable extrapolation can be obtained even by using very large values of
$a$, where small statistical errors can be obtained with
much less computational effort.
Also, the FN energies are basically independent of the shape
of the trial wave function already for a rather simple Jastrow with 1-body and 2-body terms,
implying that the ``locality error'' becomes negligible
in the variational formulation even for not-so-accurate trial wave functions. This consideration applies also
to the DMC variational energies, since the zero-lattice-space zero-time-step limits are equivalent.
\begin{figure}[t]
\includegraphics[width=1.0\columnwidth]{lrdmc_fit.pdf}
\caption[]{FN LRDMC energies as a function of lattice-space $a$ for the Oxirane (C$_2$H$_4$O) molecule, obtained using
three types of Jastrow factors. The fitting curves include a quadratic and quartic term, namely
$f(a) = E_0 + b a^2 + c a^4$.
The finite lattice space error improves dramatically with a better wavefunction.
For the simple 2-body Jastrow, $b \approx -0.15$, while for the most accurate Jastrow factor
$b' \approx -0.03$.
The ratio of the variances of the two trial wave functions is roughly equal to $b/b'$,
in agreement with Eq.~(\ref{smalla}).
}
\label{fig:lrdmc}
\end{figure}
\subsection{Relative efficiency}
In all the methods presented here, there is an extra computational cost
per Monte Carlo step with respect to
the standard DMC with LA since an extra step is needed in order to sample
correctly the Green's function related to the non-local pseudopotentials.
However, we have seen that, in \emph{all} the variational methods,
the non-local pseudopotential operator
will displace only one electron a time since the non-local pseudopotential gives a
one-body contribution to the Hamiltonian. This means that, in order to
update all the quantities needed by the simulation as the wave function ratios,
$V_{{\bf R}',{\bf R}}$, the gradients and Laplacian terms, one can exploit
the Sherman-Morrison algebra, which scales as $N^2$. For instance, to
update the non-local term $V_{{\bf R}',{\bf R}}$, as well as the ${\cal K}^a$ in the LRDMC,
one employs the same algebra as the one used to update the gradient (i.e. the drift term)
in the standard DMC with importance sampling. After a single particle move,
the cost to \emph{fully} update $V_{{\bf R}',{\bf R}}$ scales as $\mathrm{n_\mathrm{off}}~ N^2$
where $\mathrm{n_\mathrm{off}}$ is the number of non-local mesh points per electron ($\mathrm{n_\mathrm{off}}=\mathrm{n_\mathrm{quad}}$
in DMC with non-local pseudopotentials and $\mathrm{n_\mathrm{off}}=\mathrm{n_\mathrm{quad}}+6$ in LRDMC with
non-local pseudopotentials and a single lattice-space in the Laplacian).
In the size-consistent DMC (both ``version 1'' and ``version 2''), the pseudopotential move
has to be performed $N$ times in a single time-step $\tau$, so the overall cost per time-step coming
from the pseudopotential operator is $\eta ~ \mathrm{n_\mathrm{off}} ~ N^3$, where $\eta$ is the acceptance ratio of
the non-local part. Since $\mathrm{n_\mathrm{off}} \approx 20$,
and $\eta \approx 0.1$ at convergence (see Fig.~\ref{fig:jwalkalize}),
it is clear that the DMC ``version 1'' will be only a prefactor $\approx 2$ slower than the
standard DMC with LA. The ``version 2'' might be slightly slower than the ``version 1''
since it requires the calculation
of the local energy after each single-particle move, but again the difference will be just a prefactor.
The LRDMC approach is the slowest because the total number of operations
in a cycle with $N$ single-electron updates of the local energy
takes $ (10+\mathrm{n_\mathrm{quad}}) /( 4 +\mathrm{n_\mathrm{quad}})\le 2.5$ more operations
(the worst case is for full-core calculations when $\mathrm{n_\mathrm{quad}}=0$).
Moreover, there is also an additional slowing down compared to the
DMC ``version 1'' approach because, in the latter case,
all operations involving the local energy can be done at the end of a cycle
and cast in a very efficient form using matrix-matrix multiplications
of size $\sim N$. These operations, for large $N >\simeq 1000$,
can be much more efficient than
single-electron matrix updates (by a factor ranging from 2 to 20, depending
on the computer hardware and software). At present, it is
difficult to estimate how much slower LRDMC will be on a particular machine~\cite{efficiency_LRDMC},
also considering that further algorithmic and software developments are expected
in the near future, which should allow faster updates.
However, even though LRDMC is certainly slower, it has the advantage of a much smoother lattice-space extrapolation as discussed above.
\section{Conclusions}
In conclusion, we have introduced important developments in the
DMC and LRDMC methods in the context of electronic structure simulations
with non-local pseudopotentials.
We have explained how to modify the
DMC variational formulation for non-local potentials of
Ref.~\cite{C06} in order to make it size-consistent.
We have shown that, for large systems, the original algorithm~\cite{C06}
will depart from the usual localization approximation
only for small time-steps, making the zero time-step extrapolation
possibly problematic. Instead, the two DMC algorithms presented here, based
on a more accurate Trotter break-up for the non-local operator and
a better branching factor, have a smaller and size-consistent time-step error.
The DMC version 1, which features a single-particle representation
of the non-local operator and a branching factor
symmetric with respect to the application of the diffusion operator,
is straightforward to implement in the existing codes.
The DMC version 2 is closer to the LRDMC spirit, since
the non-local part is further split in $\tau/N$ factor always
acting on the all-electron configuration, and the branching factor
is accumulated after every single-particle move.
The latter version can give an even better time-step error
(order $O(\tau/N)$ in the non-local part),
particularly for relatively poor wave functions.
In general, it is slightly more time consuming than
the version 1, since it requires the evaluation of the full non local matrix
after every single-particle move.
We have made significant progress also in the LRDMC approach.
In the present formulation, it is no longer
necessary to use two lattice meshes to randomize the electron position,
but a single lattice space $a$ is sufficient, provided the orientation of the
Cartesian coordinates of the discretized Laplacian is changed randomly
during the diffusion process.
We have defined a better lattice regularization of the Hamiltonian in order
to have always a potential bounded from below, with a cutoff depending on $a$.
This leads to a well defined and size-consistent lattice-space extrapolation since, in the
$a \rightarrow 0$ limit, we recover the variational expectation value of the
continuous Hamiltonian with a lattice space error whose leading term is quadratic in $a$.
Moreover, we showed that the prefactor of the $a^2$ term vanishes quadratically in
$|\Psi_0 - \Psi_T|$.
Therefore, for good wave functions,
the extrapolation to the $a\to 0$ limit is particularly
rapid and smooth with a computational effort $\propto 1/a^2$.
The DMC error appears instead to be less correlated to the
quality of the guiding function and may display a turn-down behavior for small time-steps (observed
here and elsewhere~\cite{needs-water}), which makes the time-step extrapolation
much harder than in the LRDMC lattice-space approach.
Regarding the computational cost, the LRDMC approach is slower but the overall efficiency is comparable
to the two variational and size-consistent DMC formulations presented here since LRDMC
allows one to work with large values of $a$ due to the robust extrapolation to the zero lattice-space limit.
\acknowledgments
CF acknowledges the support from the Stichting Nationale Computerfaciliteiten
(NCF-NWO) for the use of the SARA supercomputer facilities. MC thanks the
computational support provided by the NCSA of the University of Illinois at Urbana-Champaign. SS and SM acknowledge support from CINECA and COFIN'07.
|
\section{Introduction}
High-precision measurements of the cosmic microwave background (CMB) anisotropies and the agreement of these observations with our theoretical expectations has elevated cosmology to a precision science. The observed anisotropies have been measured to follow isotropic, Gaussian statistics with an almost scale-invariant power-spectrum. However, as observations have improved, anomalies which might indicate deviations from the theoretical expectations have increased in significance. Prosaic explanations for the observed anomalies may yet be found as our understanding of the systematics of the observations improves, but it is also of interest to explore whether these anomalies point us towards exotic and unexpected physics.
One such anomaly is the apparent breakdown of statistical isotropy that has been reported in the CMB fluctuations at the largest observable scales \cite{Copi:2003kt,deOliveiraCosta:2003pu,Eriksen:2003db,Schwarz:2004gk,Land:2005ad} measured by the Wilkinson Microwave Anisotropy Probe (WMAP). Two major observations\footnote{A quadrupolar anisotropy of the inferred primordial power spectrum has also been reported \cite{Groeneboom:2008fz, Hanson:2009gu,Groeneboom:2009cb}, which is aligned with the ecliptic. However, this signal shows significant variations between different WMAP differencing assemblies at the same frequency \cite{Hanson:2009gu}, pointing to an experimental systematic which has not been taken into account in these analyses.} that are suggestive of this statistical anisotropy refer to particular special directions in the sky. The first is the planarity of the quadrupole and octopole and their mutual alignment; this plane is also roughly orthogonal to the CMB dipole direction \cite{Land:2005ad,Copi:2003kt,deOliveiraCosta:2003pu}. The second is an asymmetry in the amplitude of the power spectrum between the northern and southern ecliptic hemispheres \cite{Eriksen:2003db} which has also been observed in Cosmic Background Explorer (COBE) data but at a lower significance \cite{Eriksen:2003db}.
With the evidence for these anomalies increasing as the data improves \cite{Land:2006bn,Hansen:2008ym} many studies have proposed modifications of early-universe physics in order to generate a violation of statistical isotropy (see Refs.~\cite{Ackerman:2007nb,Erickcek:2008sm,Erickcek:2008jp,Erickcek:2009at} for a small selection). There have also been discussions of how a violation of statistical isotropy would affect other observations assuming that its source is primordial \cite{Dvorkin:2007jp,Hirata:2009ar,Frommert:2009qw, Groeneboom:2009cb}, enabling consistency tests of those hypotheses.
However, soon after these anomalies were discovered, it was noticed that the directions associated with them corresponded to special directions within the local universe: four of the planes associated with the quadrupole and octopole are orthogonal to the ecliptic, the remaining octopole plane is orthogonal to the Galactic plane, and the hemispherical asymmetry is aligned with the ecliptic. The heuristic connection between these CMB anomalies and our local environment has generated only a handful of quantitative attempts to connect the two, including possible foregrounds associated with the heliosphere \cite{Frisch:2007kz}, the kinetic Sunyaev-Zel'dovich effect of free electrons in the Galactic disk \cite{Hajian:2007xi,Waelkens:2007wn}, the thermal Sunyaev-Zel'dovich effect of the local universe \cite{Suto1996, Abramo:2006hs}, the Rees-Sciama effect of the local superclusters \cite{Maturi:2007xr}, and the Integrated Sachs-Wolfe effect of the low-redshift universe \cite{Francis:2009pt}. This is an area which deserves greater attention, since a very local signal can affect the largest observable scales.
In this paper, we explore another possible local origin of the anomalies by investigating how the scattering of CMB photons by free electrons diffusely distributed within the Milky Way halo through the kinetic Sunyaev-Zel'dovich (kSZ) effect may introduce {\it an anisotropic contamination} of the primordial signal and help explain the origin of the anomalies and their alignment with special directions in the local universe. The effects of the kSZ from the Galactic halo on the CMB has been recently discussed in Ref.~\cite{Birnboim:2009ne}. However Ref.~\cite{Birnboim:2009ne} did not explore how the kSZ may impact these anomalies and instead concentrated on whether the local kSZ may produce a foreground that must be subtracted in order to extract the primordial signal.
The prediction of an extended gaseous halo within the Milky Way is quite robust. Models of galaxy formation for halos with masses $M \gtrsim 10^{12} M_{\odot}$ predict infalling gas is shock-heated to the virial temperature (around $10^{6}-10^{7}$ K for the Milky Way) and remains in hydrostatic equilibrium until it is able to cool and condense to form stars \cite{Keres:2004cq,Birnboim:2003xa}. For a Milky Way-sized halo it is expected that the baryonic mass fraction follows the cosmic average, $f \sim 0.1$. By subtracting the total baryonic mass in stars and gas in the Galaxy, a reasonable estimate of the baryonic mass in an extended, hot, gaseous halo is $\sim 5 \times 10^{10}\ M_{\odot}$\cite{Birnboim:2009ne}. Observations of OVII and OVIII X-ray absorption have also indicated the existence of an extended hot gaseous halo \cite{2007ApJ...669..990B} associated with the Milky Way. Furthermore, observations of pulsars yield estimates for the column density of free electrons and lead to a free-electron fraction within the Milky Way halo which is consistent with a hot gaseous halo with a column density of $\sim 10^{21}\ {\rm cm}^{-2}$ \cite{Taylor:1993my}. Given that the Milky Way is moving relative to the CMB with a velocity $v/c \sim 10^{-2}$, the local kSZ may produce a signal around 1 $\mu$K.
In \S~\ref{sec:2} we point out that any anisotropic optical depth to Thomson scattering off local electrons, coupled through the kSZ effect with the dipole of the local electrons with respect to the CMB, induces an anisotropic imprint with a black-body spectrum. Its alignment is determined by the dipole direction and the anisotropy of the distribution. Given that the kSZ is expected to be subdominant to the cosmological signal, in \S~\ref{sec:impact} we point out the disproportionate impact on anomaly statistics of small signals. We make no judgement on the usefulness or otherwise of such {\sl a posteriori} anomaly statistics (for a post-WMAP 7 year discussion of this point, see Ref.~\cite{Bennett:2010jb}). Instead, we take at face value anomaly statistics which have been invoked in a large body of literature, and study the impact of the kSZ signal on their significance. In \S~\ref{sec:doppler_dip} we derive the coupling of an anisotropic electron screen to the Doppler dipole, and consider several physical models of the form of the anisotropic electron distribution in \S~\ref{sec:models}. We compute the impact of these halo kSZ models on anomaly statistics in \S~\ref{sec:stats}, and discuss the results in \S~\ref{sec:discuss}.
\section{Effect of local kSZ on CMB anisotropies}\label{sec:2}
A simple way to understand how an anisotropic optical depth leads to additional anisotropy in the CMB is to note that along a given line of sight, $\hat{\vect{n}}$, with optical depth $\tau(\hat{{\vect{n}}})$, the fraction of photons which scatter out of the line of sight is given by
\begin{equation}
\left[1-e^{-\tau({\hat{\bf n}})}\right]\left[\bar{T}+ \Delta T(\hat{{\vect{n}}})\right],
\end{equation}
where $\Delta T(\hat{{\vect{n}}})$ is the temperature anisotropy along that line of sight. In addition, there are photons that scatter in to the line of sight isotropically from every other part of the sky; thus, they contribute the mean temperature, $\bar{T}$, and the observed temperature anisotropy is \cite{Haiman:1999me}
\begin{eqnarray}
T_{\rm obs} &=& (\bar{T} + \Delta T) - \left[1-e^{-\tau(\hat{\bf n})}\right]\left[\bar{T} + \Delta T\right] \nonumber \\
& & + \bar{T} \left[1-e^{-\tau(\hat{\bf n})}\right] \nonumber \\
&=& \bar{T} + \Delta T(\hat{{\vect{n}}}) e^{-\tau(\hat{\bf n})}.
\end{eqnarray}
From this simple calculation we can see that an anisotropic optical depth will couple with any inherent temperature anisotropies and thereby modulate the observed anisotropy.
A more rigorous way to derive the same result is to consider the Boltzmann equation which dictates the evolution of the photon distribution function. In cosmology the application of the Boltzmann equation leads to the usual evolution equations for the photon distribution function within a perturbed Friedmann-Roberson-Walker (FRW) universe \cite{Dodelson:2003ft,Challinor:2009tp}. It can also be used to describe the evolution of the photon distribution function as the photons pass through the Milky Way's halo. Denoting perturbations to the photon temperature by $\bar{T} \Theta(t,\vect{x},\hat{{\vect{n}}})$ the Boltzmann equation gives
\begin{eqnarray}
\frac{d \Theta}{dt} = \frac{\partial \Theta}{\partial t} - \hat{{\vect{n}}} \cdot \vect{\nabla} \Theta \hspace{1.8in} \label{eq:Boltzmann1} \\
= \sigma_T n_e\bigg(- \Theta + \frac{3 }{16 \pi} \int d^2 \hat{{\vect{m}}} \ \Theta(\hat{{\vect{m}}}) \left[ 1+ (\hat{{\vect{n}}} \cdot \hat{{\vect{m}}})^2\right] \nonumber \\
- \hat{{\vect{n}}} \cdot \vect{v}_b/c\bigg), \nonumber
\end{eqnarray}
where the first term describes scattering out of the line of sight, the second term is due to scattering into the line of sight, the last term is the Doppler shift due to the bulk motion of the electrons (with velocity $\vect{v}_b$ in the CMB rest-frame). The Thomson scattering cross-section is $\sigma_T = 6.65 \times 10^{-25}\ {\rm cm^2}$.
Any optical depth contributed by free electrons within the Milky Way halo will be much smaller than unity so we may solve Eq.~(\ref{eq:Boltzmann1}) perturbatively. The zeroth order solution follows $d\Theta^{(0)}/dt=0$. We then find the first order correction to this
\begin{equation}
\frac{\partial \tilde{\Theta}^{(1)}(\vect{k})}{\partial t} - i k \mu \tilde{\Theta}^{(1)} (\vect{k}) = \int d^3 x\ e^{i {\bf k} \cdot {\bf x}} \sigma_T n_e(\vect{x}) S^{(0)}(\vect{x}, \hat{{\vect{n}}}),
\label{eq:Boltzmann2}
\end{equation}
where $\mu = \hat{{\vect{n}}} \cdot \hat{{\vect{k}}}$ and $S^{(0)}$ is the term in parentheses in Eq.~(\ref{eq:Boltzmann1}) evaluated for the zeroth-order solution. In the case of the CMB this term is approximately given by $S^{(0)} = -\hat{{\vect{n}}} \cdot \vect{v}_b$ since $v_b/c = 6 \times 10^{-3}$ \cite{Kogut:1993ag} and $\tilde{\Theta}^{(0)} \sim 10^{-5}$. As can be verified by substitution, the solution to Eq.~(\ref{eq:Boltzmann2}) is then given by
\begin{equation}
\Theta^{(1)}(\hat{{\vect{n}}}) = -\hat{{\vect{n}}}\cdot \vect{v}_b\ \sigma_T \int_0^{\infty} ds\ n_e(s \hat{{\vect{n}}}),
\end{equation}
where the integral extends along the line of sight. The observed temperature pattern is then given by
\begin{eqnarray}
\Theta^{\rm obs} &=& \Theta^{\rm cos} - \hat{{\vect{n}}} \cdot \vect{v}_b\ \sigma_T \int_0^{\infty} ds\ n_e(s \hat{{\vect{n}}}) \nonumber \\
&=& \Theta^{\rm cos} - \left[\hat{{\vect{n}}} \cdot \vect{v}_b\right] \tau(\hat{{\vect{n}}}),
\label{eq:obs_T}
\end{eqnarray}
where we have explicitly written $\Theta^{(0)} = \Theta^{\rm cos}$ and $n_e$ is the number density of electrons within the Milky Way halo.
Defining $\mathcal{C}(\hat{{\vect{n}}}) \equiv \sigma_T \int_0^{\infty} ds \ n_e(s \hat{{\vect{n}}})$, the amplitude of the second term in Eq.~(\ref{eq:obs_T}) can be written
\begin{eqnarray}
\hat{{\vect{n}}}\cdot \vect{v}_b\ \sigma_T \int_0^{\infty} ds\ n_e(s \hat{{\vect{n}}}) \hspace{1.5in} \nonumber \\
= 1.33\times 10^{-6} \frac{\bar{\mathcal{C}}}{10^{21}\ {\rm cm^{-2}}}\frac{\mathcal{C}(\hat{{\vect{n}}})}{\bar{\mathcal{C}}}\frac{\hat{{\vect{n}}} \cdot \vect{v}_b}{600\ {\rm km/s}},
\label{eq:angle_avg}
\end{eqnarray}
where we have defined the angle averaged $\bar{\mathcal{C}} \equiv \int d^2 \hat{n}\ \mathcal{C}(\hat{n})/(4\pi)$.
Without any reference to a specific model for the optical depth, it is clear that the kSZ is capable of producing any modulation of the CMB that we may observe. Letting $\tau(\hat{{\vect{n}}})$ denote the optical depth as a function of position on the sky, assuming some underlying primordial temperature anisotropies then the optical depth is determined by
\begin{equation}
\tau(\hat{{\vect{n}}}) = \frac{\Theta^{\rm cos}-\Theta^{\rm obs}}{\hat{{\vect{n}}} \cdot \vect{v}_b}.
\end{equation}
\section{Disproportionate impact of small signals on anomaly statistics} \label{sec:impact}
While, in principle, this effect can generate the observed large-angle isotropy anomalies, in practice (given the expected column density of free electrons and the dipole velocity of the CMB), estimates of the kSZ are at the $1$ $\mu$K level (compared to the $\pm 20$ $\mu$K of the quadrupole). Given that, at most, the kSZ can produce $\sim$ 10\% modulations of the primary CMB, it would appear as though its effect on the observed anomalies would be negligible. However, as we now discuss, even a 10\% modulation can have a large effect on the inferred statistical significance of these anomalies. Consider the ``angular momentum dispersion'' statistic \cite{deOliveiraCosta:2003pu, Gordon:2005ai, Dvorkin:2007jp} for quantifying the planarity of multipole $\ell$:
\begin{equation}
L^2_{\ell}(\hat{{\vect{n}}}) = \frac{\sum_{m= -\ell}^{\ell} m^2 |a^{\rm obs}_{\ell m}|^2}{\ell^2 \sum_{m=-\ell}^{\ell} |a^{\rm obs}_{\ell m}|^2}\ ,
\end{equation}
which, given the observed realization on the sky, is maximized in some direction $\hat{{\vect{n}}}'$. The transformation between the general frame and the maximizing frame is given by
\begin{equation}
a_{\ell m'} = \sum_{m=-\ell}^{\ell} D^\ell_{m m'} (\tilde\phi, -\theta, -\phi)\ a_{\ell m}\ ,
\end{equation}
where $D$ is the Wigner matrix corresponding to the appropriate rotation between the frames, and $\tilde\phi$ can take any value. The statistics given by the maximum values of $L^2_2$, $L^2_3$ and $L^2_{23}=L^2_2+L^2_3$ have been used in the literature to quantify the planarity of the quadrupole and octopole, and also to capture their mutual alignment, respectively.
Table~\ref{tab:dispersionstats} shows the values and $p$-values for these statistics for the WMAP 7 year Internal Linear Combination (ILC7) foreground-cleaned map \cite{Jarosik:2010iu}, with and without subtracting the kinematic quadrupole (KQ). The $p$-values are computed from 10,000 isotropic realizations. The kinematic quadrupole induced by the dipole anisotropy due to our motion relative to the CMB is at the level of $-1$ $\mu$K $\rightarrow$ $+3$ $\mu$K, whereas the CMB quadrupole is at the $-22$ $\mu$K $\rightarrow$ $+16$ $\mu$K level. The Table shows that the subtraction of the very subdominant KQ signal leads to a factor of $\sim 2.7$ decrease in the $p$-value of the most anomalous statistic, $L^2_{23}$. This effect is even more dramatic for the alternative WMAP 5 year Tegmark-Oliveira-Costa-Hamilton (TOH5) foreground-cleaned map \cite{deOliveiraCosta:2003pu}, where the maximum value of $L^2_{23}$ goes from $0.93$ to $0.96$ upon subtracting the KQ. The $p$-value corresponding to the latter is $0.0029$, making the statistic a factor of $5.5$ more unlikely under the null hypothesis of isotropy. This dramatic impact of subtracting the small KQ signal was also noticed in the WMAP first year data by Ref.~\cite{Copietal2006}, and is also present in the WMAP 5 year ILC (ILC5) map \cite{Hinshaw:2008kr}. For purposes of comparison, the $p$-value for the ILC5-KQ $L^2_{23}=0.95$ is $0.0062$.
\begin{table}[!ht]
\caption{Values and $p$-values (computed from 10,000 isotropic realizations) for angular momentum dispersion statistics $L^2_2$, $L^2_3$ and $L^2_{23}$ (see text) for the WMAP 7 year Internal Linear Combination (ILC7) map, with or without subtracting the kinematic quadrupole (KQ).}
\begin{center}
\begin{tabular}{lcccc}
\hline
Statistic & ILC7 & ILC7 & ILC7-KQ & ILC7-KQ \\
& maximum & $p$-value & maximum & $p$-value \\
\hline
$L^2_2$ & $0.943$ & 0.66 & $0.983$ & 0.38\\
$L^2_3$ & $0.919$ & 0.17 & $0.926$ & 0.15 \\
$L^2_{23}$ & $0.931$ & 0.015 & $0.953$ & 0.0055 \\
\hline\\
\end{tabular}
\end{center}
\label{tab:dispersionstats}
\end{table}
Given these observations, it is very important to quantify the impact of the kSZ effect of the Galactic halo for two reasons. First, the velocity of the Galactic barycenter with respect to the CMB points in the direction $(l,b) = (266.5^\circ, 29.1^\circ)$ \cite{COBEDipole1993}, which is quite close to the Solar CMB dipole direction, $(263.8^\circ, 48.2^\circ)$ and thus relevant for the orientations picked out by the large-angle isotropy anomalies (see Fig.~\ref{fig:directions}). Secondly, measurements of the column density of free electrons in the Galactic halo are not very precise and come primarily from theoretical arguments; observational limits are very difficult to obtain and the uncertainties on constraints are large \cite{Bregman:2009qh}. Thus, in the following calculations, we will consider a number of plausible physical models of the geometry of the halo free electron optical depth, but allow the magnitude of the effect to float somewhat, roughly at the level of the KQ amplitude, while keeping it greatly subdominant to the CMB quadrupole/octopole amplitudes.
\begin{figure}[!ht]
\includegraphics[scale=0.5]{ilc_directions.pdf}
\caption{The WMAP 5 year quadrupole and octopole, with the ecliptic plane (solid line), the Solar system dipole (SD), and the Galactic barycenter dipole (GD).}
\label{fig:directions}
\end{figure}
\section{Coupling to Doppler dipole \label{sec:doppler_dip}}
After processing through an anisotropic Thomson scattering screen, an anisotropic temperature field $\Theta$ gets screened as $\Theta(\hat{{\vect{n}}}) \rightarrow \Theta(\hat{{\vect{n}}})\exp(-\tau(\hat{{\vect{n}}}))$, as we have seen. Consider that the electron distribution consists of an isotropic part $\bar\tau$ and a (small) anisotropy $\tau^{\rm scr}(\hat{{\vect{n}}})$. This can be expanded as $e^{-\tau(\hat{{\vect{n}}})} \simeq e^{-\bar\tau}(1-\tau^{\rm scr}(\hat{{\vect{n}}}))$, leading to an observed temperature anisotropy $\Theta^{\rm obs}(\hat{{\vect{n}}}) = e^{-\bar\tau}\left[1-\tau^{\rm scr}(\hat{{\vect{n}}})\right] \Theta(\hat{{\vect{n}}})$, with the screened component of the temperature field $T^{\rm scr}(\hat{{\vect{n}}}) = e^{-\bar\tau}\tau^{\rm scr}(\hat{{\vect{n}}})\Theta(\hat{{\vect{n}}})$.
Let us expand the anisotropic part of the local optical depth in spherical harmonics,
\begin{equation}
\tau^{\rm scr}(\hat{{\vect{n}}}) = \sum_{\ell_1\geq 1} \sum_{m_1} \tau_{\ell_1 m_1} Y_{\ell_1 m_1} (\hat{{\vect{n}}}),
\end{equation}
which we will couple to the Doppler dipole of the CMB,
\begin{equation}
T^{\rm dip}(\hat{{\vect{n}}}) = A_{\rm d} Y_{1 0} (\hat{{\vect{n}}}).
\end{equation}
The screened temperature field after passing through the anisotropic $\tau^{\rm scr}$, assuming this is much smaller than the isotropic part of the screening optical depth, is then given by
\begin{equation}
T^{\rm scr}(\hat{{\vect{n}}}) \simeq e^{-\bar\tau} \tau^{\rm scr}(\hat{{\vect{n}}}) T^{\rm dip} (\hat{{\vect{n}}})
\end{equation}
Expanding $T^{\rm scr}$ in spherical hamonics, we obtain
\begin{eqnarray}
T^{\rm scr}_{\ell m} &=& e^{-\bar\tau} \int d^2 \hat{{\vect{n}}}\ \tau^{\rm scr}(\hat{{\vect{n}}}) T^{\rm dip}(\hat{{\vect{n}}}) Y^*_{\ell m} (\hat{{\vect{n}}}), \nonumber \\
&=& e^{-\bar\tau} (-1)^m \int d^2 \hat{{\vect{n}}}\ \tau^{\rm scr}(\hat{{\vect{n}}}) T^{\rm dip}(\hat{{\vect{n}}}) Y^*_{\ell m} (\hat{{\vect{n}}}), \nonumber \\
&=& e^{-\bar\tau} (-1)^m A_{\rm d} \sum_{\ell_1 m_1} \tau_{\ell_1 m_1} \times \nonumber \\
& & \int d^2 \hat{{\vect{n}}}\ Y_{\ell_1 m_1} (\hat{{\vect{n}}}) Y_{1 0} (\hat{{\vect{n}}}) Y^*_{\ell m} (\hat{{\vect{n}}}) .
\end{eqnarray}
The right hand side of this equation can be expressed in terms of Wigner 3$j$ symbols using the Gaunt integral
\begin{eqnarray}
\int d^2 \hat{{\vect{n}}}\ Y_{\ell_1 m_1} (\hat{{\vect{n}}}) Y_{\ell_2 m_2} (\hat{{\vect{n}}}) Y_{\ell_3 m_3} (\hat{{\vect{n}}}) &=& \nonumber \\
\sqrt{\frac{(2\ell_1+1)(2\ell_2+1)(2\ell_3+1)}{4\pi}} \times \nonumber \\
\left( \begin{array}{ccc}
\ell_1 & \ell_2 & \ell_3 \\
0 & 0 & 0 \end{array} \right)
\left(\begin{array}{ccc}
\ell_1 & \ell_2 & \ell_3 \\
m_1 & m_2 & m_3 \end{array} \right),
\end{eqnarray}
leading to
\begin{eqnarray}
\label{eq:couple}
T^{\rm scr}_{\ell m} &=& e^{-\bar\tau} (-1)^m A_{\rm d} \sum_{\ell_1 m_1} \tau_{\ell_1 m_1}
\times \nonumber \\
& & \sqrt{\frac{3(2\ell_1+1)(2\ell+1)}{4\pi}} \left( \begin{array}{ccc}
\ell_1 & 1 & \ell \\
0 & 0 & 0 \end{array} \right)
\left(\begin{array}{ccc}
\ell_1 & 1 & \ell \\
m_1 & 0 & -m \end{array} \right), \nonumber \\
&=& e^{-\bar\tau} (-1)^m A_{\rm d} \sum_{\ell_1} \tau_{\ell_1 m}
\sqrt{\frac{3(2\ell_1+1)(2\ell+1)}{4\pi}} \times \nonumber \\
& & \left( \begin{array}{ccc}
\ell_1 & 1 & \ell \\
0 & 0 & 0 \end{array} \right)
\left(\begin{array}{ccc}
\ell_1 & 1 & \ell \\
m & 0 & -m \end{array} \right),
\end{eqnarray}
where the second expression has made use of the symmetry property of the Gaunt integral that $m_1 + m_2 + m_3 = 0$. The first 3$j$ symbol in Eq.~(\ref{eq:couple}) enforces the triangle condition $\ell_1=\ell \pm 1$, so finally we obtain
\begin{equation}
\label{eq:tscreen}
T^{\rm scr}_{\ell m} = e^{-\bar\tau} A_{\rm d} \sqrt{\frac{3}{4\pi}} \left[ C_{\ell-1, m} \tau_{\ell-1, m} + C_{\ell+1, m} \tau_{\ell+1, m} \right],
\end{equation}
where
\begin{eqnarray}
C_{\ell-1, m} &=& \sqrt{\frac{\ell^2 - m^2}{4\ell^2-1}}\\
C_{\ell+1, m} &=& \sqrt{\frac{(1+\ell)^2 - m^2}{(1+2\ell)(3+2\ell)}} .
\end{eqnarray}
Remember that in these expressions, the $\tau_{\ell'' m''}$ starts at $\ell''=1$. So, $T_{0m} \propto C_{1m}\tau_{1m}$, $T_{1m} \propto C_{2m}\tau_{2m}$, $T_{2m} \propto (C_{1m}\tau_{1m} + C_{3m}\tau_{3m})$, and so on.
\section{Models for anisotropic free electron distributions} \label{sec:models}
In this section we consider how an extended gaseous partially ionized halo couples the CMB to the Doppler dipole through the kSZ effect, first gaining physical intuition using an analytic toy model of an anisotropic optical depth distribution, and then considering physical models.
\subsection{Simple geometries for the electron distribution}
Let us develop a feel for some simple, plausible electron distributions. Consider $\tau(\theta,\phi)$, a smooth function defined on the unit sphere, with $(0 \leq \theta \leq \pi, 0 \leq \phi \leq 2\pi)$, which we want to expand in terms of spherical harmonics as
\begin{equation}
\tau(\theta,\phi) = \sum_{\ell=0}^\infty \sum_{m=-\ell}^\ell a_{\ell m} Y_{\ell m}(\theta, \phi) .
\end{equation}
The spherical harmonic coefficients for $\tau$ are given as usual by
\begin{equation}
a_{\ell m} = \int_0^{2\pi} \int_0^\pi d\phi\ d\theta \sin \theta\ \tau(\theta, \phi) Y^*_{\ell m} (\theta, \phi).
\end{equation}
A sphere only has the non-zero coefficient $a_{00}$ so is not useful for our purpose. Consider the ellipsoid centered on the origin and aligned with the coordinate axes,
\begin{equation}
\left(\frac{x}{a}\right)^2 + \left(\frac{y}{a}\right)^2 + \left(\frac{z}{c}\right)^2 = 1,
\end{equation}
where as usual, $x= \tau \sin \theta \cos \phi$, $y=\tau \sin\theta \sin\phi$, $z=\tau \cos \theta$. Thus for this distribution,
\begin{equation}
\tau(\theta,\phi) = ac\left( a^2 \cos^2 \theta + c^2 \sin^2 \theta \right)^{-1/2}.
\end{equation}
This function only depends on $\theta$, so the only non-zero spherical harmonic coefficients have $m=0$. Further, since $\tau(\theta,\phi)$ is even in $\theta$, only $a_{\ell 0}$ where $\ell$ is even are non-zero. This is interesting because it says that if the anisotropy in $\tau$ is axisymmetric in the dipole frame, $T^{\rm scr} = T^{\rm scr}_{\ell 0}$ where only $\ell=$ {\em odd} multipole couplings would survive. While in the real world, any local free electron distribution is unlikely to be anisotropic in a axisymmetric way in the CMB dipole frame, we can see that in general there will be an asymmetry between odd and even multipoles in the power spectrum of the screening field. This analytic argument thus explains the ``sawtooth" pattern of the power spectrum of the kSZ effect of the Galactic disk found by Ref.~\cite{Hajian:2007xi}.
Having developed analytic intuition for the form of the expected screening field, we will now present several physical models for the anisotropic optical depth.
\begin{figure}[!htp]
\includegraphics[height=20pc]{alms.pdf}
\caption{The power spectrum of the full screening field corresponding the four models for the free-electron distribution in the halo. The solid black curve corresponds to the generalized NFW halo with a core radius of $r_0 = 300$ kpc ({\sf SphSym}). The short-dashed blue and red curves correspond to a triaxial halo (with $e_b = e_c = 0.5$) aligned with ({\sf Triax}) and perpendicular to ({\sf TriaxRot}) the angular momentum of the disk, respectively. The long-dashed black curve corresponds to the triaxial model for the Milky Way halo proposed in Ref.~\cite{Law:2009yq} in order to explain the tidal tails of the Sgr dSph ({\sf TriaxLMJ}). As described in Eq.~(\ref{eq:angle_avg}) the amplitude of the signal depends linearly on the angle averaged column-density and is shown here for the fiducial case $\bar{\mathcal{C}} = 10^{21}{\rm cm}^{-2}$. The discussion in Sec.~\ref{sec:doppler_dip} gives an analytic explanation for the sawtooth pattern found in power-spectra generated by these free-electron distributions. }
\label{fig:alms}
\end{figure}
\begin{figure}[!ht]
\includegraphics[height=41pc]{maps_hires.pdf}
\caption{Sky maps of the screening field for the three cases, from top to bottom: a generalized spherical NFW halo with $a= 0.01$ ({\sf SphSym}); a generalized triaxial NFW halo with major axis perpendicular to the plane of the Galactic disk, $a=0.01$, and $e_b=e_c =0.5$ ({\sf Triax}); the same as before but with the major axis within the plane of the Galactic disk ({\sf TriaxRot}). In both cases the intermediate axis is chosen to be perpendicular to the line connecting the Solar system and the Galactic center. In the final case ({\sf TriaxLMJ}) we take the model for the Galactic halo recently proposed in Ref.~\cite{Law:2009yq} in which the minor axis is aligned with the line connecting the Solar system and Galactic center and the major axis lies within the Galactic disk.}
\label{fig:maps}
\end{figure}
\subsection{Physical models for the anisotropic optical depth} \label{sec:anisomodel}
There are two basic ways that we may imagine an anisotropic optical depth through the Galaxy. In the first case, the Galaxy is assumed to have a spherically symmetric gaseous halo and the anisotropy is a consequence of the fact that the Solar system is offset from the Galactic center. The IAU standard for the Galactocentric distance of the Sun is 8 kpc; however, this value has been recently revised upwards by $\sim 5\%$ \cite{Ghez08, Gillessen09}. We adopt the conservative Galactocentric distance 8.5 kpc to maximize the anisotropic optical depth arising from this offset. In the second case, it is also possible that the gaseous halo is triaxial, leading to an additional source of anisotropy. We take the virial radius of the Milky Way to be 300 kpc (which corresponds to an NFW halo with a mass of 1.5 $\times 10^{12} \ M_{\odot}$ and a concentration $c=12$).
\subsubsection{Spherically symmetric halo}
In order to explore the case of a spherically symmetric halo we need only specify the number density of electrons as a function of the radial distance from the Galactic center (GC). As discussed in Ref.~\cite{Birnboim:2009ne} N-body simulations which include gas dynamics generically predict a hot gaseous halo for Milky Way-like galaxies. For the spherically symmetric case we will consider a generalized NFW density profile with an inner core
\begin{equation}
n_e = \frac{n_0}{(1+r/r_0)^3}.
\end{equation}
Spherically symmetric profiles can be written in a self-similar form (where the origin is at the GC)
\begin{equation}
n_e = n_0 y_{\rm gas}(r/r_0).
\end{equation}
Using the Galactic coordinate system (where the origin coincides with the location of the Solar system and the GC lies on the positive $x$-axis) the line-of-sight density takes the form
\begin{equation}
n_e(t,\theta,\phi) = n_0 y_{\rm gas}\left(\sqrt{t^2 + a^2 - 2a t \cos(\phi) \sin(\theta)}\right),
\end{equation}
where $t \equiv r/r_0$ and $a \equiv r_{GC}/r_0$. The normalization of the density distribution is set by requiring the mean angular column density be equal to $10^{21}\ {\rm cm}^{-2}$.
The normalization for this model can be computed analytically as
\begin{equation}
\bar{\mathcal{C}} = n_0 r_0 \frac{1-a^2 + 2 a^2 \log(a)}{2(1-a^2)^2}.
\end{equation}
\subsubsection{Triaxial halos}
N-body simulations show that dark matter halos are not spherically symmetric but instead are better approximated by triaxial density profiles \cite{1988ApJ...327..507F,1991ApJ...378..496D,1992ApJ...399..405W,Jing:2002np}. This triaxiality, if reflected in the density distribution of an extended gaseous halo, will produce additional anisotropy in the optical depth. Fits to N-body simulations \cite{Jing:2002np} show that isodensity contours are well approximated by a radial coordinate
\begin{equation}
R\equiv x^2 + \frac{y^2}{1-e_b^2}+\frac{z^2}{1-e_c^2},
\end{equation}
where the $x$-axis and $z$-axis run along the major and minor principal axes, respectively; $e_b$ and $e_c$ ($e_b<e_c$) are the ellipticities of the halo isodensity contours. Cosmological N-body simulations give a nearly Gaussian distribution for $e_b$ with $\langle e_c \rangle \approx 0.8$ and $e_b \lesssim 0.7$ \cite{Jing:2002np}. We also note that it has been suggested the inclusion of gas cooling may make the dark-matter density distribution more spherical even out to the virial radius \cite{Kazantzidis:2004vu}.
As with the spherically symmetric case, we consider a generalized triaxial NFW halo with a core. As suggested by Ref.~\cite{Zentner:2005wh} in order to obtain consistency between observations of the distribution of sub-halos about the Milky Way and the $\Lambda$CDM prediction, the major axis of the triaxial halo must be aligned with the disk angular momentum. In order to explore how the relationship of the disk angular momentum and the triaxiality of the dark matter halo affects the local kSZ signal, we consider the two cases where the major axis is aligned with or perpendicular to the disk angular momentum. In both cases, the intermediate axis is along the Galactic $y$-axis\footnote{The Galactic coordinate system is centered on the location of the Solar system with the Galactic $x$-axis pointing towards the Galactic center and the Galactic $z$-axis pointing in the direction of the angular momentum of the Galactic disk.}.
We also consider the recent triaxial Milky Way halo proposed in Ref.~\cite{Law:2009yq} in order to explain the observed characteristics of the Sagittarius dwarf spheriodal (Sgr dSph). Using observations of the tidal streams of the Sgr dSph, Ref.~\cite{Law:2009yq} claim a Milky Way halo with a minor/major axis ratio $e_b \approx 0.56$ and an intermediate/major axis ratio of $e_c \approx 0.74$, with the minor axis lying along the Galactic $x$-axis and the major axis along the Galactic $y$-axis.
\section{Effect of halo kSZ on isotropy anomalies} \label{sec:stats}
Given a model of the anisotropic optical depth, we obtain a prediction for $T_{\rm scr}$ from Eqs.~(\ref{eq:couple}) or ~(\ref{eq:tscreen}). We can add that to the observed realization to obtain the actual cosmological realization, and calculate whatever anomaly statistics we want for the cosmological realization.
\subsection{Quadrupole--octopole planarity and alignment}
First, let us consider the angular momentum dispersion statistics of \S~\ref{sec:impact}. Taking the four models for the halo described in \S~\ref{sec:anisomodel}, labelled {\sf SphSym}, {\sf Triax}, {\sf TriaxRot} and {\sf TriaxLMJ} respectively, we compute the screening fields and add them to the WMAP ILC7 map after subtraction of the kinematic quadrupole. In order to estimate the amplitude of the signal at which the $p$-value of the statistic would be affected for each case, we scale up the maximum of the screening map to an amplitude (1 $\mu$K, 2 $\mu$K, 5 $\mu$K) which is much smaller than the primordial quadrupole and octopole signals, before adding to the ILC7 map. The 1 $\mu K$ signal is not large enough to make any difference to the statistics. The results for the other two cases are presented for the $L^2_{23}$ statistic in Table~\ref{tab:screeningstats}. The $L^2_2$ and $L^2_3$ statistics of the ILC7-KQ map are not anomalous before or after the kSZ correction and thus we do not repeat them here.
\begin{table}[!ht]
\caption{Values and $p$-values (computed from 10,000 isotropic realizations) for angular momentum dispersion statistic $L^2_{23}$ for the WMAP 7 year Internal Linear Combination (ILC7) map subtracting the kinematic quadrupole (KQ) after accounting for the anisotropic screening from the four models described in the text, after scaling the maximum amplitude of the screening map to 2 $\mu$K and 5 $\mu$K. We also list the corresponding value for the angle-averaged free electron column density, $\bar{\mathcal{C}}$, in units of $10^{21}\ {\rm cm}^{-2}$.}
\begin{center}
\begin{tabular}{| l | c|c|c||c|c|c|}
\hline
Model & $L^2_{23}$ & $p$-value & $\bar{\mathcal{C}}$ & $L^2_{23}$ & $p$-value & $\bar{\mathcal{C}}$\\
& (2 $\mu$K) & (2 $\mu$K) & (2 $\mu$K) & (5 $\mu$K) & (5 $\mu$K) & (5 $\mu$K) \\
\hline
{\sf SphSym} & 0.955 & 0.0047 & 12 & 0.952 & 0.0056 & 31 \\
{\sf Triax} & 0.962 & 0.0022 & 22 & 0.972 & 0.0008 & 56 \\
{\sf TriaxRot} & 0.950 & 0.0062 & 24 & 0.943 & 0.0091 & 60 \\
{\sf TriaxLMJ} & 0.956 & 0.0042 & 4 & 0.958 & 0.0036 & 11 \\
\hline
\end{tabular}
\end{center}
\label{tab:screeningstats}
\end{table}
In interpreting Table~\ref{tab:screeningstats}, the baseline $p$-value to compare with is $0.0055$ for the ILC7-KQ map with no kSZ correction. The Table shows that for the {\sf Triax} case, small signals at the $\sim 2-5$ $\mu$K level {\sl reduce} this already small $p$-value by factors of $\sim 2$--$7$! When scaled to $5$ $\mu$K, the {\sf TriaxRot} $p$-value increases somewhat by factors of $\lesssim 2$ and the {\sf TriaxLMJ} $p$-value decreases by roughly the same factor. The {\sf SphSym} screening map is unable to affect this statistic significantly at these signal levels. These tests show that a statistic such as $L^2_{23}$, which is sensitive to the orientation of the Galactic plane, can be highly sensitive to small signals oriented with this plane.
If the maximum amplitude of the screening map is between 2--5 $\mu$K, the average free electron column densities for these maps are between $4 \times 10^{21}$ -- $6 \times 10^{22}\ {\rm cm}^{-2}$. As we discuss in more detail below, though the lower end of this range is possible given current observational constraints, the upper range is unlikely.
\subsection{Hemispherical asymmetry}
We now explore the extent to which the scattering of CMB photons in an extended gaseous halo can account for the observed hemispherical asymmetry \cite{Eriksen:2003db}. The statistic by which this asymmetry is inferred assumes a dipolar modulation of an isotropic primordial CMB temperature field, and constrains the relative amplitude of this modulation. The dipolar modulation leads to a coupling between modes of order $\ell$ and $\ell\pm1$. As we saw in Sec.~\ref{sec:doppler_dip}, the kSZ also introduces such a coupling between modes, so it is particularly interesting to consider how, through the kSZ, a screening field may impact the significance of the dipole modulation amplitude statistic for constraining hemispherical asymmetry.
In order to do this, we will use the approach presented in Refs.~\cite{Gordon:2005ai, Dvorkin:2007jp}. In that work an estimator is derived which measures to what extent the data shows a dipolar modulation by looking for a non-zero coupling between multipoles of order $\ell$ and $\ell\pm1$ up to some $\ell_{\rm max}$,
\begin{equation}
\hat{w}_1^{TT} = \frac{\sum_{\ell m} \frac{f_{\ell}^{TT} R^{1 \ell}_{\ell+1,m}}{C_{\ell}^{TT} C_{\ell+1}^{TT}}(a^T_{\ell m})^* (a^T_{\ell+1 m})}{\sum_{\ell m} \frac{(f_{\ell}^{TT} R^{1 \ell}_{\ell+1,m})^2}{C^{TT}_{\ell} C^{TT}_{\ell+1 m}}},
\end{equation}
where
\begin{eqnarray}
f_{\ell}^{TT} &\equiv& C_{\ell}^{TT} + C_{\ell+1}^{TT}\\
R^{\ell_1, \ell_2}_{\ell m} &\equiv& (-1)^m \sqrt{\frac{(2 \ell+1)(2 \ell_1+1)(2 \ell_2 +1)}{4\pi}} \times \nonumber \\
& & \left(\begin{array}{ccc}\ell_1 & \ell_2 & \ell \\0 & 0 & 0\end{array}\right)\left(\begin{array}{ccc}\ell_1 & \ell_2 & \ell \\0 & m & -m\end{array}\right).
\end{eqnarray}
From the discussion in Sec.~\ref{sec:doppler_dip} it is clear that the inclusion of Thomson scattering from free electrons in an extended Galactic halo will couple multipoles of order $\ell$ to those of order $\ell-1$ and $\ell +1$ in the observed $a_{\ell m}$s.
\begin{figure}[!ht]
\includegraphics[height=20pc]{hemispherical_asymmetry.pdf}
\caption{How a screening field can affect the measurement of a hemispherical asymmetry. The grey band indicates the confidence level of the distribution at that value of $\ell_{\rm max}$ ranging from 1 $\sigma$ to 3 $\sigma$. The solid thick black curve shows the value of $\hat{w}_1$ evaluated on the ILC cleaned WMAP7 map. From top to bottom we show the resulting hemispherical asymmetry for {\sf TriaxLMJ}, {\sf SphSym}, {\sf TriaxRot}, and {\sf Triax} as a function of $\ell_{\rm max}$. This Figure demonstrates how the orientation of the triaxiality of the halo greatly affects its ability to impact the inferred amplitude of a hemispherical asymmetry.}
\label{fig:hemiasym}
\end{figure}
To check the effect of the kSZ on the significance of the hemispherical asymmetry, we evaluate the estimator $\hat{w}_1^{TT}$ on the inferred cosmological signal given by $T^{\rm cos}_{\ell m} = T^{\rm obs}_{\ell m} + T^{\rm scr}_{\ell m}(\bar{\mathcal{C}})$ for various values of the angle-averaged electron column density $\bar{\mathcal{C}}$. As in the previous section, the four extended halos considered here are the spherically symmetric case ({\sf SphSym}), and the three triaxial halos with various orientations ({\sf Triax}, {\sf TriaxRot}, {\sf TriaxLMJ}).
We show the value of the estimator $\hat{w}_1$ evaluated on the WMAP ILC7 map in Fig.~\ref{fig:hemiasym}. Because of the scale dependence of the power spectra of the screening fields, seen in Fig.~\ref{fig:alms}, we plot the value of $\hat{w}_1$ as a function of $\ell_{\rm max}$ in Fig.~\ref{fig:hemiasym}.
The grey band indicates the confidence level of the distribution at that value of $\ell_{\rm max}$ ranging from 1 $\sigma$ to 3 $\sigma$. The Figure shows the value of $\hat{w}_1$ for (from top to bottom) for {\sf TriaxLMJ}, {\sf SphSym}, {\sf TriaxRot}, and {\sf Triax}. It is clear that the orientation of any triaxiality has a significant impact on the inferred amplitude of the hemispherical asymmetry. In particular, the {\sf TriaxLMJ} orientation leads to an increase in the inferred amplitude. This can be seen in Fig.~\ref{fig:maps} since the inclusion of the {\sf TriaxLMJ} screening field adds power in the northern Galactic hemisphere and subtracts it from the southern hemisphere. For the other cases, the inferred amplitude decreases.
Since the anisotropies induced by a local kSZ are at the level of $10^{-5}\ \mu{\rm K}$ for $\ell \gtrsim 10$ (see Fig.~\ref{fig:alms}), in order for the screening field to have a significant impact on the inferred amplitude of $\hat{w}_1$, we must have $\bar{\mathcal{C}}/(10^{21}\ {\rm cm^{-2}}) \sim 10^4$. A free-electron column density of this magnitude is ruled out by observations of OVII and OVIII absorption \cite{2007ApJ...669..990B} and pulsar observations towards the Large Magellanic Cloud (LMC) \cite{Taylor:1993my}. Therefore, we find that a local kSZ is unlikely to provide a plausible explanation for a hemispherical asymmetry.
\section{Discussion} \label{sec:discuss}
Although measurements of the CMB have generally confirmed our current understanding of the formation and evolution of the universe, there are several anomalies which appear to be at odds with our standard picture.
These anomalies could suggest that there may be some process which violates the statistical isotropy of the CMB fluctuations on large scales. Although attempts have been made to modify the physics of the early universe in order to explain these anomalies, there are also strong reasons to look for explanations in the local universe. First, the directionality of these anomalies seems to closely coincide with both the Solar CMB dipole as well as the velocity of the Galactic barycenter relative to the CMB rest-frame. Second, given that the anomalies are on large angular scales, any non-primordial and causal explanation must be local in origin.
Here we have explored how a local kSZ signal (due to the motion of any extended hot gaseous halo associated with the Milky Way relative to the CMB) may affect the significance of several anomalies observed in the CMB anisotropies. Both theoretical and observational considerations indicate the presence of a hot gaseous halo with an extent of several tens of kiloparsecs \cite{Bregman:2009qh, Birnboim:2009ne}. Anisotropies in the optical depth through this gaseous halo can be due to the offset of the Solar system from the Galactic center, as well as any triaxiality in its distribution. We computed the kSZ signal from several plausible physical models for the shape and orientation of the halo, and studied their impact on the observed CMB sky, in the form of the WMAP ILC7 map.
We considered how a local kSZ may affect the observed planarity of the CMB quadrupole and octopole moments as well as their relative alignment, motivated by the fact that kSZ from a triaxial halo can naturally relate to the special directions that appear to be associated with the $\ell=2$ and $3$ multipole moments. Surprisingly, we found that relatively small changes in the observed amplitude of these moments ($\sim 10\%$) to account for a kSZ signal can reduce the already tiny $p$-values for these anomalies of up to factors of 2--7. The corresponding free electron column density needed to affect the planarity/alignment statistics at this level is $4$ -- $60 \times 10^{21}\ {\rm cm}^{-2}$.
A local kSZ signal couples multipole moments of order $\ell$ to those of order $\ell \pm 1$, so it is natural to consider how such a signal would affect any inferred hemispherical asymmetry in the CMB anisotropies in the form of a dipolar modulation, which gives rise to the same coupling. Using a statistic first derived in Ref.~\cite{Dvorkin:2007jp}, we found that a local kSZ signal can have a significant effect on the inferred amplitude of a dipolar modulation of the primary CMB anisotropies. In this case, the corresponding free electron column density would need to be $\sim10^{25}\ {\rm cm}^{-2}$.
Theoretical and observational constraints on the free electron fraction in a extended hot gaseous halo associated with the Milky Way place a fairly strict bound on the column density of $< 10^{21}\ {\rm cm}^{-2}$. The most precise constraints come from observations of the dispersion measure to individual pulsars \cite{Taylor:1993my} which find a column density in free electrons to the LMC of $\sim 3 \times 10^{20}\ {\rm cm}^{-2}$. Given that the LMC is $\sim 50$ kpc from the Galactic center and that the scale-radius of the Milky Way halo is $\sim 300$ kpc, these observations indicate that the total optical depth may be as large as $10^{21}\ {\rm cm}^{-2}$. In addition to this, observations of local ($z=0$) OVII and OVIII absorption towards several quasars indicates the presence of an extended hot gaseous halo around the Milky Way \cite{Bregman:2009qh}. The inferred free electron column density is quite uncertain given assumptions about the metallicity of the gas (Solar abundance was assumed to convert the observations to an electron density; sub-Solar values would increase the inferred density), as well as a model dependence coming from the assumed profile of the gas. Finally, theoretical considerations imply the existence of a hot extended gaseous halo with a fractional mass of the order of the cosmic baryon fraction, $f\sim 0.1$. Therefore, for a Milky Way sized halo ($M \sim 1.5 \times 10^{12}\ M_{\odot}$) with a scale radius of 300 kpc, we would expect a column density no larger than $10^{21}\ {\rm cm}^{-2}$.
Given that, for a local kSZ signal to have a significant effect on the CMB anisotropies, we must have a free electron column density $> 10^{21} \ {\rm cm}^{-2}$, but that both theoretical and observational considerations place the limit at $\lesssim 10^{21} \ {\rm cm}^{-2}$, it is unlikely that a local kSZ signal can explain any of the CMB anomalies considered here. Even given the uncertainties, the kSZ signal can at best only be at the lowest amplitude needed to affect the anomaly statistics.
Finally, we note that contamination from kSZ in the Solar system can also, in principle, provide an anisotropic contamination of the CMB sky at large angles. The geometry of the heliopause is coincidentally aligned with the CMB dipole \cite{Frisch:2007kz}, providing a motivation for looking there for a explanation for the special directional properties of the large-angle isotropy anomalies. However, in practice, the optical depth of free electrons in the Solar system is more than 7 orders of magnitude below what is required to produce the necessary signal.
\acknowledgments{HVP is supported by Marie Curie grant MIRG-CT-2007-203314 from the European Commission, and by STFC and the Leverhulme Trust. TLS is supported by the Berkeley Center for Cosmological Physics. We thank the Aspen Center for Physics, where this work was initiated, for hospitality. We are grateful to Hsiao-Wen Chen, Priscilla Frisch, Shirley Ho, Ed Jenkins, and Chris Thom for invaluable information about observational constraints on the free electron content of the local universe. We thank Yuval Birnboim for useful comments on some preliminary results. HVP thanks Anthony Challinor, Daniel Mortlock, and Andrew Pontzen for interesting discussions on related topics.}
\input{bibliography.bbl}
\end{document} |
\section{Introduction}
Recently, there has been considerable interest in developing optical analogues of black holes in table-top devices~\cite{Philbin-2008, Zhang-2009,Narimanov-2009,Cui-2009}. The electromagnetic parameters of these devices are designed in such a way as to mimic effects arising in the presence of black holes. Using this approach, an artificial event horizon has been implemented by means of a moving medium~\cite{Philbin-2008}, an inhomogeneous refractive index profile was proposed to mimic celestial mechanics~\cite{Zhang-2009}, and a broadband omnidirectional absorber has been developed~\cite{Narimanov-2009,Cui-2009}. Although the equivalence between gravitational effects and optical media was suggested by several authors~\cite{Balazs-1957,Plebanski-1960,Felice-1971}, it has recently been conceived from a different point of view~\cite{Ward-1996} and has become a tool to design optical devices within the framework of transformation optics~\cite{Pendry-2006,Leonhardt-2006,Leonhardt-2006-2, Leonhardt-2009}.
Transformation optics provides a new way of looking at the interaction between light and matter. This recently developed method in electromagnetism is based on the analogy between the macroscopic Maxwell's equations in complex materials and the free-space Maxwell's equations on the background of an arbitrary metric. It describes how a general topological deformation of reality can be implemented using materials with nontrivial constitutive parameters. The invisibility cloak, which bends the electromagnetic reality, i.e., the electromagnetic energy flow, around a hole---which therefore turns invisible---, is probably the most famous example of this technique~\cite{Pendry-2006, Leonhardt-2006,Cai-2007,Valentine-2009}. Moreover, the cloaking principle extends beyond electromagnetism and can be applied in acoustics~\cite{Cummer-2008}, hydrodynamics~\cite{Farhat-2008}, and even quantum mechanics~\cite{Zhang-2008}.
Up to now, most applications of transformation optics have been related to the spatial manipulation of optical beams and the design of imaging systems \cite{Rahm2-2008, Kwon-2008, Jacob-2006}. These devices are generated through spatial coordinate transformations and do not consider curved geometries. Transformation optics, however, applies to general four-dimensional metrics, and thus extends beyond the scope of existing applications. In this paper, we show how a four-dimensional metric that plays a role in general relativity---the Robertson-Walker metric---can be implemented directly with dielectrics and we demonstrate theoretically that the cosmological redshift can be reproduced inside a dielectric structure of finite size with time-varying electromagnetic properties.
\section{Electromagnetic analogue of the cosmological redshift}
The cosmological redshift of an electromagnetic wave traveling from one place to another is an effect that originates from the expansion of the universe. Such an expanding background can be modeled by the Robertson-Walker metric,
\begin{equation}
ds^2 = -c^2\mathrm{d}t^2 + a^2(t)\left[\frac{\mathrm{d}r^2}{1-\kappa r^2} + r^2\, d\Omega^2\right],
\end{equation}
where $a(t)$ is a real, dimensionless scale factor, related to the well-known Hubble parameter by $H = \dot{a}/a$, and $\kappa$ represents the spatial curvature of the metric. An electromagnetic wave with frequency $\omega_\mathrm{em}$ emitted at instance $t = t_\mathrm{em}$ will be observed at a different frequency $\omega_\mathrm{obs}$ at time $t = t_\mathrm{obs}$. These two frequencies are related by (see, e.g., Ref.~\cite{Carroll-2003})
\begin{equation}
\frac{\omega_\mathrm{obs}}{\omega_\mathrm{em}} = \frac{a(t_\mathrm{em})}{a(t_\mathrm{obs})}.
\label{Eq:CosmologicalRedshift}
\end{equation}
In an expanding universe, the frequency of electromagnetic waves is redshifted.
\begin{figure}
\begin{center}
\includegraphics[height=3.1cm]{./ElementarySolutionFinal}\label{Fig:RobertsonWalkerSolutionZ}
\end{center}
\caption{A graphical representation of a traveling wave solution inside a medium that is the electromagnetic analogue of the Robertson-Walker metric. (a) The spatial variation at constant time $t=1$ and (b) the temporal variation of this solution for fixed location $z = 1$, when we modulate the permittivity and permeability with $a(t) = 1 + t$.}
\label{Fig:RobertsonWalkerSolution}
\end{figure}
To construct an electromagnetic analogue of this effect, we proceed by calculating the material properties of a dielectric medium using transformation optics~\cite{Pendry-2006,Leonhardt-2006-2, Leonhardt-2009}. It is straightforward to show that the Robertson-Walker metric without spatial curvature ($\kappa=0$) can be translated into an isotropic, homogeneous dielectric with constitutive parameters
\begin{equation}
\epsilon^x_{\phantom{x}x} = \mu^x_{\phantom{x}x} = a(t),\;\;
\epsilon^y_{\phantom{y}y} = \mu^y_{\phantom{y}y} = a(t),\;\;
\epsilon^z_{\phantom{z}z} = \mu^z_{\phantom{z}z} = a(t).\label{Eq:RobertsonWalkerMaterials}
\end{equation}
The traveling wave solutions inside such a medium are found by solving Maxwell's equations in the absence of free sources and currents.
In combination with Eq.~(\ref{Eq:RobertsonWalkerMaterials}), Maxwell's equations combine into a generalized wave equation,
\begin{equation}
\Delta\mathbf{E} - \frac{1}{c^2}\frac{\partial}{\partial t}\left(a(t)\frac{\partial}{\partial t}(a(t) \mathbf{E})\right) = 0,
\end{equation}
where $c$ is the speed of light in vacuum. Its solutions are waves with time-dependent amplitude propagating along an arbitrary direction labeled by the unit vector $\mathbf{1}_\mathrm{k}$:
\begin{equation}
\label{Eq:Superposition}
\mathbf{E}(\mathbf{r},t) = \frac{1}{a(t)}\, G\left(\mathbf{r}\cdot\mathbf{1}_\mathrm{k}- c \int_{t_0}^t \frac{\mathrm{d}\tilde{t}}{a(\tilde{t})}\right)\mathbf{1_\mathrm{E}},
\end{equation}
where $G(x)$ is an arbitrary differentiable function. To gain some insight in these traveling waves, we consider a harmonic function $G(x)=\cos(x)$ in Fig.~\ref{Fig:RobertsonWalkerSolution} and we plot the temporal and spatial variation of Eq.~(\ref{Eq:Superposition}). We see that the wave has a well-defined wavelength, but that its frequency is changing in time. A straightforward calculation of the instantaneous frequency $\omega_\mathrm{inst}(t) = -\partial_t(\mathbf{k}\cdot\mathbf{r} - |\mathbf{k}| c \int_{t_0}^t \mathrm{d}\tilde{t}/a(\tilde{t}))$ reveals that the frequencies at two different times $t_1$ and $t_2$ satisfy $\omega_\mathrm{inst}(t_2)/\omega_\mathrm{inst}(t_1) = a(t_1)/a(t_2)$, in agreement with the redshift formula of Eq.~(\ref{Eq:CosmologicalRedshift}).
\section{A frequency converter of finite size}
As shown above, a traveling wave in an infinite Robertson-Walker medium is subject to an equivalent of the cosmological redshift. This does not guarantee, however, that we can use this medium to alter the frequency of an incident wave, since any practical application requires materials of finite extent. Furthermore, as was pointed out in Ref.~\cite{Rahm-2008}, the boundary conditions play a crucial role in a transformation-optical device of finite extent. We therefore examine in this section whether a slab of the Robertson-Walker medium with thickness $L$ (see Fig.~\ref{Fig:RobertsonWalkerSetup}) would suit as a frequency converter. To simplify notation, we introduce the function
\begin{equation}
\label{Eq:FrequencyTunerFt}
f(t) = \int_{t_0}^t\frac{\mathrm{d}\tilde{t}}{a(\tilde{t})}.
\end{equation}
\begin{figure}
\begin{center}
\includegraphics[height=5.5cm]{./SetupFinal}
\end{center}
\caption{A setup where we implement a finite Robertson-Walker device. At the leftmost boundary a monochromatic wave of frequency $\omega_0$ and amplitude $A_0$ illuminates the device. We modulate the permittivity and permeability as $\epsilon(t) = \mu(t) = a(t) = 1 + t$. We show that the output undergoes a frequency shift according to the cosmological redshift formula.}
\label{Fig:RobertsonWalkerSetup}
\end{figure}
We illuminate the dielectric from the left in region $(\mathrm{I})$ with a monochromatic plane wave of frequency $\omega_0 = |\mathbf{k}_0| c$, and we want to calculate the wave that is emitted in region $(\mathrm{III})$.
In general, each interface would give rise to transmitted and reflected waves, and one would explicitly impose
continuity of the tangential components of both $\mathbf{E}$ and $\mathbf{H}$. However, since the impedance $\eta = (\mu/\epsilon)^{1/2}$ is the same on both sides of each interface, we can anticipate that there are no
reflected waves. We therefore restrict our attention to the electric field $\mathbf{E}$, consider purely right-moving waves and impose continuity at each interface. In regions $(\mathrm{I})$, $(\mathrm{II})$, and $(\mathrm{III})$, respectively, the electric field can be written as
\begin{eqnarray}
\label{Eq:FieldInRegion2}
\mathbf{E}_\mathrm{I}(z,t) &=& A_0\,\mathrm{e}^{\mathrm{i}k_0 \left(z-ct\right)}\mathbf{1_\mathrm{E}},\\
\mathbf{E}_\mathrm{II}(z,t) &=& \frac{G\left(z-cf(t)\right)}{a(t)}\mathbf{1_\mathrm{E}},\\
\label{Eq:FieldInRegion3}
\mathbf{E}_\mathrm{III}(z,t) &=& H\left(z-c\, t\right)\mathbf{1_\mathrm{E}},
\end{eqnarray}
where $G(x)$ and $H(x)$ are differentiable functions. The continuity of the tangential component of the electric field at $z = 0$ yields
\begin{equation}
A_0\mathrm{e}^{-\mathrm{i} k_0 c t} = \frac{G\left(-cf(t)\right)}{a(t)}.
\end{equation}
This equation can be used to evaluate $G(x)$ for an arbitrary value of its argument $x$:
\begin{equation}
\label{Eq:WaveInsideComponent}
G(x) = a\left(f^{-1}\left(-\frac{x}{c}\right)\right) A_0\mathrm{e}^{-\mathrm{i} k_0 c f^{-1}\left(-\frac{x}{c}\right)},
\end{equation}
where we have introduced the inverse function $f^{-1}$. In Fig.~\ref{Fig:ztDiagram}, we plot the evolution of the wavefronts inside the material as a function of space and time. We can now combine Eq.~(\ref{Eq:FieldInRegion2}) with Eq.~(\ref{Eq:WaveInsideComponent}) to determine the electric field at the rightmost boundary ($z=L$) of the device:
\begin{equation}
\mathbf{E_\mathrm{II}}(L,t) = \frac{a\left(f^{-1}\left(f(t)-\frac{L}{c}\right)\right)}{a(t)}A_0\mathrm{e}^{-\mathrm{i} k_0 c f^{-1}\left(f(t)-\frac{L}{c}\right)}\mathbf{1_\mathrm{E}}.
\end{equation}
To retrieve a general expression for the electric field in region ($\mathrm{III}$), we have to express the continuity of the electric field at the rightmost boundary and we ultimately find that
\begin{equation}
\mathbf{E_\mathrm{III}}(z,t) = \mathbf{E_\mathrm{II}}\left(L,t-\frac{z-L}{c}\right).
\end{equation}
\begin{figure}
\begin{center}
\includegraphics[height=5.0cm]{./ZTDiagramFinal}
\end{center}
\caption{The wave fronts inside the frequency converter as a function of space and time when the leftmost boundary is illuminated by a monochromatic wave. In this diagram, we have simulated a linear evolution of the scale factor $a(t) = t$.}
\label{Fig:ztDiagram}
\end{figure}
The instantaneous frequency of the electromagnetic wave emitted in the vacuum region $(\mathrm{III})$ behind the device ($z=L$) can be written as
\begin{equation}
\label{Eq:FrequencyShift}
\omega_\mathrm{out} = \omega_0 \frac{a\left(f^{-1}\left(f(t)-\frac{L}{c}\right)\right)}{a(t)}.
\end{equation}
This equation shows that the proposed design from Fig.~\ref{Fig:RobertsonWalkerSetup} indeed executes a frequency shift. Since the velocity of the wavefronts $v = c/\sqrt{\epsilon\mu}=c/a(t)$, we know that $L = \int_{t_1}^{t_2} c/a(t)\, \mathrm{d}t = c\left(f(t_2)-f(t_1)\right)$, where $t_1$ and $t_2$ indicate the time of incidence and departure of a wavefront. If we are observing at time $t$ at the rightmost boundary, then $f^{-1}\left(f(t)-L/c\right)$ corresponds to the time when the wavefront was at the leftmost boundary. Hence, we retrieve the cosmological redshift formula, ${\omega_\mathrm{inst}(t_2)}/{\omega_\mathrm{inst}(t_1)} = {a(t_1)}/{a(t_2)}$. The reader should also note that the amplitude is altered with a similar factor, so that the device might also be used as an optical amplifier. It will often be beneficial to create a constant frequency shift, i.e., to generate a monochromatic wave in region (III); in this case, the right hand side of Eq.~(\ref{Eq:FrequencyShift}) has to be constant.
By renaming $f(t)= x$, this condition immediately implies that $a(f^{-1}(x)) = \exp(\alpha x) p(x)$, where $\alpha$ is an arbitrary (real) constant and $p(x)$ is a periodic function with period $L/c$.
Using Eq. (\ref{Eq:FrequencyTunerFt}), we observe that $a(f^{-1}(x))=(f^{-1})^\prime (x)$ and infer that the condition is satisfied when $f^{-1}(x)=\exp(\alpha x) q(x) - \beta$ with $\beta$ constant and $q(x)$ periodic with period $L/c$. By inverting
this relation to obtain $f(t)$ and differentiating with respect to $t$, we can calculate those $a(t)$ which generate a monochromatic wave in region (III). For instance, the special case $q(x)=1$ leads to $f(t)=\ln(t+\beta)/\alpha$ and thus to the linear profile $a(t)=\alpha(t+\beta)$.
The resulting frequency shift is then given by
\begin{equation}
\frac{\omega_\mathrm{out}}{\omega_0} = \mathrm{e}^{-\alpha L/c}.
\label{Eq:DeviceFormula}
\end{equation}
\section{Conclusion and discussion}
We note that the output signal has all its energy at the desired frequency without the creation of sidebands. The conversion is thus characterized by a very high efficiency. The proposed device achieves a frequency shift using \emph{linear} materials, so that the superposition principle remains valid and an arbitrary wavepacket can be frequency-shifted. This frequency shift is possible due to the time evolution of its parameters, which renders it non-stationary \cite{Budko-2009}.
Equation~(\ref{Eq:DeviceFormula}) enables us to calculate the characteristics of a device that generates the desired frequency shift of a monochromatic wave. Let us suppose, for instance, that we want to shift an incident frequency of $f_0 = \unit{100}{\tera\hertz}$ by \unit{10}{\giga\hertz} in a material of thickness $L = 2 \lambda_0 = \unit{6}{\micro\meter}$. We then find that the modulation rate of the refractive index must equal $\alpha \approx \unit{5}{\giga\hertz}$. Since the corresponding wavelength of this modulating signal is much larger than the dimensions of the device, we can generate the permittivity variation by low-frequency electro-optic modulation. The most difficult part is the time evolution of the permeability. This can be achieved by introducing an array of split-ring resonators \cite{Smith-2004, Soukoulis-2006} in the electro-optic material.
The frequency-dependent permeability of an array of split rings can be modeled by
\begin{equation}
\mu(\omega)=1+\frac{F\omega^2}{\omega_\mathrm{LC}^2-\omega^2},
\end{equation}
where $F$ is the filling factor and $\omega_\mathrm{LC}$ is the resonance frequency, which is inversely proportional to the square root of the material's permittivity $\epsilon_\mathrm{C}$ inside the gap of the split rings. Operating far below resonance, we can approximate the permeability by
$\mu(\omega)\approx1+F\omega^2 c^2 l^2 \epsilon_\mathrm{C}/(dw),$
with $l$, $d$, $w$ the geometrical parameters of the split rings \cite{Soukoulis-2006}. By properly choosing the electro-optic material inside the gap of the split rings, we can thus vary the permeability and permittivity according to Eq.~(\ref{Eq:RobertsonWalkerMaterials}) with electro-optics to establish a frequency shift determined by Eq.~(\ref{Eq:DeviceFormula}).
\section*{Acknowledgments}
Work at the VUB was supported by the Belgian Science Policy Office (Grant No. IAP VI/10 Photonics@be \& Grant No. IAP VI/11), by the FWO-Vlaanderen (Project G011410N, fellowships), and by the Research Council (OZR) of the VUB. Work at Ames Laboratory was supported by the Department of Energy (Basic Energy Sciences) under Contract No. DE-AC02-07CH11358. P.~T. acknowledges the Belgian American Educational Foundation for financial support.
\end{document}
|
\section*{Introduction}
In the classical case the tensor product theorem is formulated for either totally disconnected groups or Lie groups and the proof by D. Flath \cite{Flath}, which is found in the books until this day, uses different methods for each case.
The tensor product theorem is used in the theory of automorphic forms, to prove that every irreducible admissible representation of an adele group $G({\mathbb A})$ is an infinite product of local representations.
In this note we give a unified proof which is simpler and applies to all locally compact groups.
The proof essentially contains no new ideas. We only streamline and unify the existing proofs by using a more universal concept of Hecke algebras.
This proof will also appear in the upcoming book by the author on Automorphic Forms.
\section{Preliminaries}
For a locally compact group $G$ we denote by $\hat G$ the \emph{unitary dual}, i.e., the set of isomorphy classes of irreducible unitary representations of $G$.
For a representation $(\pi,V_\pi)$ of the compact group $K$ and an irreducible representation $(\tau,V_\tau)$ of $K$
we define the \emph{$\tau$-isotype} as the sum of all subrepresentations of $\pi$, which are isomorphic to $\tau$.
If $\pi$ is unitary, then $V_\pi$ is the direct sum of its isotpyes
$
V_\pi\=\bigoplus_{\tau\in\hat K}V_\pi(\tau).
$
For a given representation $\tau\in\hat K$ let
$$
e_\tau(k)\=(\dim \tau)\ \ol{\operatorname{tr}(\tau(k))},\qquad k\in K.
$$
\begin{lemma}\label{7.5.10}
The function $e_\tau$ is an idempotent in the convolution algebra $C(K)$, i.e., one has $e_\tau *e_\tau=e_\tau$.
Let $(\pi,V_\pi)$ be a representation of $K$. Then the map
$$
P_\tau\=\int_Ke_\tau(k)\pi(k)\,dk
$$
is a projection onto the isotype $V_\pi(\tau)$.
Here we have normalized the Haar measure on $K$ so that ${\rm vol}(K)=1$.
If $\pi$ is unitary, then $P_\tau$ is the orthogonal projection onto $V_\pi(\tau)$.
If $\gamma$ is another irreducible unitary representation of $K$ which is not isomorphic to $\tau$, then
$
e_\tau*e_\gamma\= 0.
$
\end{lemma}
{\bf Proof: }
This is a direct consequence of the Peter-Weyl Theorem \cite{HA2}.
\phantom{}\ensuremath{\hfill\square}
Let $F\subset\hat K$ be a finite subset.
Then the function
$$
e_F\ \stackrel{\mbox{\rm\tiny def}}{=}\ \sum_{\tau\in F}e_\tau
$$
is an idempotent in $C(K)$, because
$$
e_F*e_F\=\sum_{\tau,\gamma\in F}e_\tau*e_\gamma\=\sum_{\tau\in F} e_\tau\= e_F.
$$
\section{Main Theorem}
Let $G$ be a locally-compact group and $K\subset G$ a compact subgroup.
A unitary representation $(\pi,V_\pi)$ of $G$ is called \emph{$K$-admissible}, if the representation $\pi|_K$ decomposes with finite multiplicities, i.e., if for every $\tau\in\widehat K$ the multiplicity
$$
[\pi:\tau]\=\dim\operatorname{Hom}_K(V_\tau,V_\pi)\=\dim\operatorname{Hom}_K(V_\pi,V_\tau)
$$
is finite.
The representation $\pi$ is $K$-admissible if and only if every $K$-isotype $V_\pi(\tau)$ is finite-dimensional.
We denote by \emph{$\widehat G_K$} the set of all isomorphy classes of irreducible unitary $G$-representations which are $K$-admissible.
A representation $\pi$ of $G$ is called \emph{admissible}, if there is a compact subgroup $K$ such that $\pi$ is $K$-admissible.
We write $\widehat G_{\rm adm}$ for the set of isomorphy classes of irreducible unitary admissible representations of $G$.
\begin{theorem}[Main Theorem]\label{4.5.9}
Let $G$ and $H$ be locally compact groups.
For any two $\pi\in\widehat G_{\rm adm}$ and $\eta\in\widehat H_{\rm adm}$ the tensor product $\pi\otimes\eta$ is in $\widehat{G\times H}_{\rm adm}$ and the tensor map $(\pi,\eta)\mapsto \pi\otimes\eta$ is a bijection
$$
\widehat G_{\rm adm}\times\widehat H_{\rm adm}\ \tto\cong\ \widehat{G\times H}_{\rm adm}.
$$
More precisely, given two compact subgroups $K\subset G$ and $L\subset H$,
the tensor map map is a bijection $\widehat G_K\times\widehat H_L\to\widehat{G\times H}_{K\times L}$.
\end{theorem}
The injectivity of the tensor map follows from the next lemma.
\begin{lemma}\label{7.6.3}
Let $G$ be a locally compact group and let $(\eta,V_\eta)$ be an irriducible unitary representation.
Let $W$ be a Hilbert space.
Then every closed $G$-stable subspace of $V_\eta\otimes W$ is of the form $V_\eta\otimes W_1$ for a closed subspace $W_1$ of $W$.
Here the tensor product refers to the tensor product in the category of Hilbert spaces, i.e., the Hilbert-space completion of the algebraic tensor product.
In particular, it follows that for two locally compact groups $G,H$ and irreducible unitary representations $\pi,\tau$ of $G$ and $H$ respectively, the representation $\pi\otimes\tau$ of $G\times H$ is irreducible and that $\pi\otimes\tau\ \cong\ \pi'\otimes\tau'$ implies $\pi\cong\pi'$ and $\tau\cong\tau'$.
\end{lemma}
{\bf Proof: }
Let $U\subset V_\eta\otimes W$ be a closed, $G$-stable subspace.
Then the orthogonal projection $P$ with image $U$ is a bounded operator on $V_\eta\otimes W$ which commutes with all $\eta(g)$, $g\in G$.
We show that every such operator $T$ is of the form $1\otimes S$ for an operator $S\in{\cal B}(W)$.
So let $T$ be a bounded operator on $V_\eta\otimes W$ with
$$
T(\eta(g)\otimes 1)\=(\eta(g)\otimes 1)T
$$
for every $g\in G$.
Let $(e_i)_{i\in I}$ be an orthonormal basis of $W$ and for $j\in I$ let $P_j$ be the orthogonal projection onto the subspace ${\mathbb C} e_j$.
For $i,j\in I$ consider the operator $T_{i,j}$ given as composition:
$$
V_\eta\to V_\eta\otimes e_i\tto T V_\eta\otimes W\tto{1\otimes P_j} V_\eta\otimes e_j\to V_\eta.
$$
This operator is bounded and commutes with the $G$-action.
By the Lemma of Schur there exists a complex number
$a_{i,j}$ with $T_{i,j}=a_{i,j}{\rm Id}$.
Since $\sum_jP_j={\rm Id}_W$, we get
$$
T(v\otimes e_i)\=\sum_ja_{i,j}v\otimes e_j\=v\otimes S(e_i),
$$
where $S(e_i)=\sum_ja_{i,j}e_j$ and the sum converges in $W$.
As $T$ is continuous, $S$ extends to a unique continuous operator $S:W\to W$, satisfying
$$
T(v\otimes w)\=v\otimes S(w)
$$
for all $v\in V_\eta$, $w\in W$.
This means $T=1\otimes S$, as claimed.
We can apply this to the orthogonal projection $P$ onto the invariant subspace $U$, which therefore is of the form $P=1\otimes P_1$.
Then $P_1$ is again an orthogonal projection.
Let $W_1$ be the image of $P_1$, then $U= V_\eta\otimes W_1$.
We next show the irreducibility of $\pi\otimes\tau$.
The first part implies that a closed, $G\times H$-stable subspace $U$ has to be of the form $U=V_\pi\otimes U_\tau$ as well as of the form $U=U_\pi\otimes V_\tau$ with closed subspaces $U_\pi\subset V_\pi$ and $U_\tau\subset V_\tau$.
These subspaces have to be invariant themselves, hence the full spaces.
Finally let $\pi\otimes\tau\cong\pi'\otimes\tau'$,
so there is a unitary isomorphism $T:V_\pi\otimes V_\tau\to V_{\pi'}\otimes V_{\tau'}$ commuting with the $G\times H$-operation.
Choose $w\in V_\tau$ with $\norm w=1$.
Then $V_\pi\otimes w$ is an irreducible $G$-subrepresentation, therefore there is $w'\in V_{\tau'}$ mit $\norm{w'}=1$, such that $T(V_\pi\otimes w)=V_{\pi'}\otimes w'$.
The map
$$
V_\pi\to V_\pi\otimes w\tto T V_{\pi'}\otimes w'\to V_{\pi'}
$$
is a unitary $G$-isomorphism, so we get $\pi\cong\pi'$ and analogously $\tau\cong\tau'$.
\phantom{}\ensuremath{\hfill\square}
\section{The Hecke algebra}
Let $G$ be a locally compact group.
The set $C_c(G)$ of all continuous functions of compact support is a complex algebra under the convolution product
$$
f*g(x)\=\int_Gf(y)g(y^{-1}x)\,dy\=\int_Gf(xy)g(y^{-1})\,dy,\quad f,g\in C_c(G).
$$
Let $K\subset G$ be a compact subgroup.
For $\alpha,\beta\in C(K)$, there is the $K$-convolution
$$
\alpha *\beta(k)\=\int_K\alpha(l)\beta(l^{-1}k)\,dl
$$
making $C(K)$ a ${\mathbb C}$-algebra as well.
We can also define a convolution between functions on $G$ and $K$ as follows.
For $\alpha\in C(K)$ and $f\in C_c(G)$ we define
$$
\alpha*f(x)\=\int_K\alpha(k)f(k^{-1}x)\,dk\quad \text{and}\quad f*\alpha(x)\=\int_Kf(xk)\alpha(k^{-1})\,dk.
$$
We normalize the Haar-measure of the compact group $K$ so that ${\rm vol}(K)=1$.
If $K$ is not a nullset in $G$, we insist that the Haar measure of $G$ be normalized also to give $K$ the volume $1$.
Hence in this case the Haar measures of $G$ and $K$ coincide on $K$.
\begin{lemma}
With this normalization the convolution products are compatible in the sense that
$$
f*(g*h)\=(f*g)*h.
$$
holds for all combinations of $f,g,h$ being functions in either $C_c(G)$ or $C(K)$.
Also we always have
$
(f*g)^*\=g^**f^*,
$
where $f^*(x)=\Delta(x^{-1})\ol{f(x^{-1})}$ and $\Delta(x)$ is the modular function of $G$.
\end{lemma}
{\bf Proof: }
A computation.\phantom{}\ensuremath{\hfill\square}
Let $G$ be a locally compact group with a compact subgroup $K$.
A function $f$ on $G$ is called \emph{$K$-finite},
if the set of all functions $x\mapsto f(k_1xk_2)$, $k_1,k_2\in K$ spans a finite-dimensional vector space.
It is a simple observation that the convolution product of two $K$-finite functions is $K$-finite.
We define the \emph{Hecke-algebra} ${\cal H}={\cal H}_{G,K}$
of the pair $(G,K)$ as the convolution algebra of all $K$-finite functions in $C_c(G)$.
\begin{lemma}\label{Lem7.5.25}
Let $G$ be alocally compact group and $K$ a compact subgroup.
\begin{enumerate}[\rm (a)]
\item Let $I_K$ be the set of all finite subsets of $\hat K$.
For $F\in I_K$ set $e_F=\sum_{\tau\in F}e_\tau$ and
$$
C_F\ \stackrel{\mbox{\rm\tiny def}}{=}\ e_F* C_c(G)*e_F\=e_F* {\cal H} *e_F.\index{$C_F$}
$$
Then $e_F$ is an idempotent in $C(K)$ and $C_F$ is a subalgebra of ${\cal H}$.
The Hecke-algebra ${\cal H}$ is the union of all these subalgebras.
If $F\subset F'$ in $I_K$, then $C_F\subset C_{F'}$.
If $F=\{\tau\}$ we also write $C_F=C_\tau$.
\item For a unitary representation $(\pi,V_\pi)$ of $G$ the space $\pi({\cal H})V_\pi$ is dense in $V_\pi$.
\item The Hecke-algebra is a *-Algebra with involution $f^*(x)=\Delta(x^{-1})\ol{f(x^{-1})}$.
If $\pi$ is a unitary representation of $G$, then $\pi$
defines a *-representation of ${\cal H}$.
For two unitary representations $\pi, \pi'$ of $G$ we have
$$
\pi\cong_G\pi'\ \Leftrightarrow\ \pi\cong_{\cal H}\pi',
$$
so, $\pi$ and $\pi'$ are unitarily equivalent if and only if they define isomorphic ${\cal H}$-modules.
\end{enumerate}
\end{lemma}
{\bf Proof: }
Part (a) is clear. We show (b).
For $h\in C(K)$ we write
$$
\pi(h)\=\int_K h(k)\pi(k)\,dk.
$$
Let $F\in I_K$ and $P_F=\pi(e_F)$.
Then $P_F^2=\pi(e_F)\pi(e_F)\=\pi(e_F*e_F)\=\pi(e_F)=P_F$.
So $P_F$ is a projection with image $V_\pi(F)=\bigoplus_{\tau\in F}V_\pi(\tau)$, which is the \emph{$F$-isotype} of $\pi$ and the kernel is $\bigoplus_{\tau\notin F}V_\pi(\tau)$.
Hence the union of all $V_\pi(F)$ with $F\in I_K$ is dense in $V_\pi$.
So it suffices to show that $\pi(C_F)V_\pi$ is dense in $V_\pi(F)$.
Let $v\in V_\pi(F)$ and let $\varepsilon>0$.
Since $\pi(e_F)$ is continuous, there is $C>0$ such that $\norm{\pi(e_F)w}\le C\norm w$ for every $w\in V_\pi$.
Since the map $G\times V_\pi\to V_\pi$; $(g,v)\mapsto \pi(g)v$ is continuous, there is a neighborhoor $U$ of the unit in $G$, such that $x\in U\ \Rightarrow\ \norm{\pi(x)v-v}<\varepsilon/C$.
Let $f\in C_c(G)$ with support in $U$ such that $f\ge 0$ and $\int_G f(x)\, dx\= 1$.
Then
$$
\norm{\pi(f)v-v}\=\norm{\int_Gf(x)(\pi(x)v-v)\,dx}
\le {\int_Gf(x)\norm{\pi(x)v-v}\,dx} <\varepsilon/C.
$$
One has $e_F*f*e_F\in C_F\subset{\cal H}$ and
$$
\norm{\pi(e_F*f*e_F)v-v}\=\norm{\pi(e_F)(\pi(f)v-v)}< \varepsilon.
$$
This shows (b).
We continue with (c).
The closedness under * is clear.
Let $\pi$ be a unitary $G$-representation, then for $f\in{\cal H}$ one has
\begin{eqnarray*}
\pi(f^*)&=&\int_G\Delta(x^{-1})\ol{f(x^{-1})}\pi(x)\,dx
\=\int_G\ol{f(x)}\pi(x^{-1})\,dx\\
&=&\int_G\ol{f(x)}\pi(x)^*\,dx\=\(\int_G{f(x)}\pi(x)\,dx\)^*\=\pi(f)^*.
\end{eqnarray*}
If $\pi$ and $\pi'$ are isomorphic as $G$-representations, then so they are as ${\cal H}$-modules.
For the converse let $T: V_\pi\to V_{\pi'}$ be a unitary
${\cal H}$-isomorphism, so $T\pi(f)=\pi'(f)T$ for every $f\in{\cal H}$.
We first remark, that the same must hold for $f\in C_c(G)$.
For this let $S=T\pi(f)-\pi'(f)T$, then for every finite subset $F$ of $\widehat K$,
$$
\pi'(e_F)S\pi(e_F)\=T\pi(\underbrace{e_F fe_F}_{\in{\cal H}})-\pi'(e_Ffe_F)T\=0.
$$
Therefore $Sv=0$ for every vector $v\in V_\pi(F)$ and since the $V_\pi(F)$ span a dense subspace of $V_\pi$, we get $S=0$, so $T\pi(f)=\pi'(f)T$ for every $f\in C_c(G)$.
Let $\varepsilon>0$ and let $x\in G$, $v\in V$.
Then, by the continuity of the representations $\pi$ and $\pi'$ there exists a neighborhood $U$ of $x\in G$ such that $\norm{\pi(u)v-\pi(x)v}<\varepsilon/2$ and $\norm{\pi'(u)Tv-\pi'(x)Tv}<\varepsilon/2$.
Let $f\in C_c(G)$ have support inside $U$ and suppose $f\ge 0$ and $\int_Gf(x)\,dx=1$.
Since $T$ is unitary, we have
\begin{multline*}
\norm{T\pi(f)v-T\pi(x)v}
\=\norm{\pi(f)v-\pi(x)v}\\
\=\norm{\int_Gf(u)(\pi(u)v-\pi(x)v)\,du}\ \le\ \int_Gf(u)\norm{\pi(u)v-\pi(x)v}\,du\ <\ \varepsilon/2
\end{multline*}
and likewise $\norm{T\pi(f)v-\pi'(x)Tv}=
\norm{\pi'(f)Tv-\pi'(x)Tv}<\varepsilon/2$.
This implies $\norm{T\pi(x)v-\pi'(x)Tv}<\varepsilon$.
Since $\varepsilon>0$ and $v\in V_\pi$ are arbitrary, we conclude $T\pi(x)=\pi'(x)T$.
\phantom{}\ensuremath{\hfill\square}
If $(\pi,V_\pi)$ is an irreducible unitary representation of $G$, then every $f\in C_c(G)$ defines a continuous linear operator $\pi(f)$ on $V_\pi$.
For $0\ne v\in V_\pi$ the space $\pi(C_c(G))v$ is a $G$-stable subspace of $V_\pi$.
As $\pi$ is irreducible, the space $\pi({\cal H})v$ is a dense subspace.
\begin{proposition}\label{4.5.11}
Let $G$ be a locally compact group and $K\subset G$ a compact subgroup.
\begin{enumerate}[\rm (a)]
\item
Let $(\pi,V_\pi)$ be an irreducible representation of $G$.
Let $F$ be a finite subset of $\hat K$.
Then the $F$-isotype $F$-Isotyp $V_\pi(F)=\bigoplus_{\tau\in F}V_\pi(\tau)$ is an irreducible $C_F$-module.
In particular, $\pi({\cal H})V_\pi$ is an irreducible ${\cal H}$-module.
\item
Let $(\eta,V_\eta)$ be a unitary representation of $G$ and $F$ a finite subset of $\hat K$.
If $M$ is a finite-dimensional irreducible $C_F$-submodule of $V_\eta(F)$, then the $G$-subrepresentation of $\pi$ generated by $M$ is irreducible.
\item
Let $\eta,\pi$ be irreducible unitary representations of $G$, which are $K$-admissible. Then
$$
\eta\cong\pi\ \Leftrightarrow\ V_\eta(F)\ \cong\ V_{\pi}(F) \text{ for every } F\in I_K,
$$
where on the right we consider isomorphy as $C_F$-modules.
\end{enumerate}
\end{proposition}
{\bf Proof: }
(a) To show irreducibility of $V_\pi(F)$, let $0\ne U\subset V_\pi(F)$ be a closed submodule.
Since $V_\pi$ is irreducible, the space $\pi(C_c(G))U$ is dense in $V_\pi$.
Let $P_F:V_\pi\to V_\pi(F)$ be the isotypical projection.
For $h\in C(K)$ we write
$$
\pi(h)\=\int_Kh(k)\pi(k)\,dk.
$$
Then we have $P_F=\pi(e_F)$.
Let $f\in C_c(G)$. Then
$$
\pi(f)\pi(e_\tau)=\pi(f*e_\tau)\quad\text{und}\quad \pi(e_\tau)\pi(f)=\pi(e_\tau *f),
$$
as a calculation shows.
The projection $P_F$ is continuous, so the space $P_F(C_c(G) U)$ is dense in $V_\pi(F)$, therefore $\ol{P_F(C_c(G)U)}=V_\pi(F)$.
Since $P_F=\pi(e_F)$ we get
\begin{eqnarray*}
V_\pi(F)&=&\ol{P_F(C_c(G)U)}\=\ol{\pi(e_F)C_c(G)U}\\
&=&\ol{\pi(e_F*C_c(G)*e_F)U}\=\ol{\pi(C_F)U}\=\ol{U}\= U,
\end{eqnarray*}
which shows that $V_\pi(F)$ is irreducible.
We now show (b).
Let $m_0\in M\smallsetminus\{ 0\}$ and let
$
\operatorname{Ann}(m_0)\=\{a\in C_F:am_0=0\}
$
be its annihilator.
Then $\operatorname{Ann}(m_0)$ is a left ideal of $C_F$ and the map $a\mapsto am_0$ is a module isomorphism
$$
C_F/\operatorname{Ann}(m_0)\to C_Fm_0.
$$
Since $M$ is finite-dimensional and irreducible it follows $M=C_Fm_0$, so $M\cong C_F/J$ with $J=\operatorname{Ann}(m_0)$.
Let $U$ be the $G$-stable closed subspace of $V_\pi$ generated by $M$.
We show $P_F(U)=M$, where $P_F=\pi(e_F)$ is the orthogonal projection onto $M$.
\emph{Claim:} Let $\ol J$ be the annihilator $\operatorname{Ann}_{C_c(G)}(m_0)$ of $m_0$ in $C_c(G)$.
Then
$$
\ol J\=\ol J e_F \oplus \operatorname{Ann}_{C_c(G)}(e_F),
$$
where $\operatorname{Ann}_{C_c(G)}(e_F)$ is the annihilator of $e_F$ in $C_c(G)$, which is the set of all $f\in C_c(G)$ with $f*e_F=0$.
Ww show tha claim.
It is clear that $\ol Je_F\subset\operatorname{Ann}_{C_c(G)}(v_0)=\ol J$. Further $v_0=e_F v_0$ and so $\operatorname{Ann}_{C_c(G)}(e_F)\subset\operatorname{Ann}_{C_c(G)}(v_0)=\ol J$.
It remains to show that $\ol J$ lies in the right hand side.
Since $e_F$ is an idempotent, we have $C_c(G)=C_c(G)e_F\oplus \operatorname{Ann}_{C_c(G)}(e_F)$, since every $f\in C_c(G)$ can be written as $f=fe_F+(f-fe_F)$ and $f-fe_F$ lies in the annihilator of $e_F$.
Further
$
\operatorname{Ann}_{C_c(G)}(e_F)\ \subset\ \operatorname{Ann}_{C_c(G)}(v_0)=\ol J
$, which shows the claim.
So we have
$$
C_c(G)v_0\cong C_c(G)/\ol J\cong C_c(G)e_F/ (\ol Je_F).
$$
This implies
$$
P_F(C_c(G)v_0)
\cong e_F C_c(G)e_F/(\ol Je_F)
\cong C_F/J\cong M.
$$
Therefore $P_F(U)=M$.
Let now $U'$ be a subrepresentation of $U$. Then $P_F(U')=0$ or $M$. In the first case $M\subset (U')^\perp$ and so $U\subset (U')^\perp$, since the space $(U')^\perp$ is a subrepresentation.
This gives $U'=0$.
In the second case $M\subset U'$ and so $U'=U$.
Indeed we conclude that $U$ is irreducible.
(c) Let $(\eta,V_\eta),(\pi,V_\pi)$ be isomorphic representations, then so are the $C_F$-modules $V_\eta(F)$ and $V_\pi(F)$.
For the converse let for every $F$ be given a $C_F$-isomorphism
$\phi_F:V_\eta(F)\to V_\pi(F)$.
According to the Lemma of Schur two isomorphisms for a given $F$ only differ by a multiple scalar.
So we can normalize all these isomorphisms in a manner that they extend each other, i.e., that we have
$\phi_F=\phi_{F'}|_{V_\eta(F)}$, if $F\subset F'$.
This means that the $\phi_F$ can be put together to a ${\cal H}$-isomorphism
$$
\phi:\eta({\cal H})V_\eta\tto\cong\pi({\cal H})V_\pi.
$$
Both sides are dense sub vector spaces.
If we can show that $\phi$ is isometric, this map extends to an isomorphism
of $G$-representations by Lemma \ref{Lem7.5.25}.
Let $0\ne v_0\in \eta({\cal H})V_\eta$ and let $w_0=\phi(v_0)$.
We can assume that $\norm{v_0}=\norm{w_0}=1$ and we claim that $\phi$ is isometric.
For this we transport both inner products to the same sode via $\phi$, so we have two inner products $\sp{.,.}_1$ and $\sp{.,.}_2$ on, say $\pi({\cal H})V_\pi$ such that $v_0$ has norm $1$ in both inner products and the operation of ${\cal H}$ is a $*$-operation in both inner products.
On the finite dimensional space $\pi(C_F) V_\pi$ these inner products must coincide by the Lemma of Schur.
This holds for every $F\in I_K$, so that $\phi$ is indeed isometric.
\phantom{}\ensuremath{\hfill\square}
Let now $G,H$ locally compact with compact subgroups $K\subset G$ and $L\subset H$.
Let $E$ be a finite subset of $\hat K$ and $F$ a finite subset of $\hat L$.
There is a natural Homomorphism
$$
\psi:C_c(G)\otimes C_c(H)\to G_c(G\times H)
$$
given by
$$
\psi(f\otimes g)(x,y)\=f(x)g(y).
$$
This induces a homomorphism
$$
C_E\otimes C_F
\to C_{E\times F}.
$$
This map is not surjective in general.
But this fact causes no harm, as we have
\begin{lemma}\label{Lem7.5.22}
If $M$ is a finite-dimensional irreducible $C_{E\times F}$-*-submodule of a unitary $G$-representation, then it is irreducible as $C_E\otimes C_F$-module.
\end{lemma}
{\bf Proof: }
We equip $C_{E\times F}\subset L^1(G\times H)$ with the topology of the $L^1$-norm.
Then $C_E\otimes C_F$ is dense in $C_{E\times F}$, as $C_c(G)$ is dense in $L^1(G)$.
The representation $\rho:C_{E\times F}\to\operatorname{End}_{\mathbb C}(M)$ is continuous, since $M$ comes from a unitary representation of $G\times H$.
So the image of $C_E\otimes C_F$ is dense in the (finite-dimensional) image of $C_{E\times F}$ in $\operatorname{End}(M)$,
so these two images agree.
\phantom{}\ensuremath{\hfill\square}
In order to show Theorem \ref{4.5.9} we only need the second assertion of the following lemma.
\begin{lemma}\label{Lem7.5.23}
\begin{enumerate}[\rm (a)]
\item Let $A$ be a ${\mathbb C}$-algebra with unit and $M$ a simple $A$-module, which is finite dimensional over ${\mathbb C}$.
Then the map $A\to\operatorname{End}_{\mathbb C}(M)$ is surjective.
\item Let $A,B$ complex algebras with units and let $R=A\otimes B$.
If $M,N$ are simple modules over $A$ and $B$ resp., which are finite-dimensional over ${\mathbb C}$, then $M\otimes N$ is a simple $R$-module and every simple $R$-module, which is finite-dimensional over ${\mathbb C}$, is of this form with uniquely determined modules $M$ and $N$.
\end{enumerate}
\end{lemma}
{\bf Proof: }
Part (a) is a well-known result of Wedderburn.
See \cite{Lang} for a proof.
We show (b). According to (a) it suffices to assume $A=\operatorname{End}_{\mathbb C}(M)$ and $B=\operatorname{End}_{\mathbb C}(N)$ for the first assertion.
The canonical map from $\operatorname{End}_{\mathbb C}(M)\otimes\operatorname{End}_{\mathbb C}(N)$ to $\operatorname{End}_{\mathbb C}(M\otimes N)$ is surjective, and so $M\otimes N$ is simple.
Let now $V$ be a given ${\mathbb C}$-finite-dimensional simple $A\otimes B$-module.
Then $V$ contains a simple $A$-module $M$ as $V$ is finite-dimensional.
Let $N=\operatorname{Hom}_A(M,V)$.
This vector space is a $B$-module in the obvious way.
Consider the map $\phi:M\otimes N\to V$ given by
$$
\phi(m\otimes \alpha)\=\alpha(m).
$$
Then $\phi\ne 0$ is an $A\otimes B$-homomorphism, hence surjective, as $V$ is simple.
We have to show that $\phi$ has zero kernel.
For this let $\alpha_1,\dots,\alpha_k$ be a basis of $N$ and $m_1,\dots,m_l$ a basis of $M$.
Let $c_{i,j}\in{\mathbb C}$ be given with $\phi\(\sum_{i,j}c_{i,j}m_i\otimes\alpha_j\)=0$.
We have to show that all coefficients $c_{i,j}$ are zero.
We have
$$
0\=\phi\(\sum_{i,j}c_{i,j}m_i\otimes\alpha_j\)\=\sum_{i,j}c_{i,j}\alpha_j(m_i)\=\sum_j\alpha_j\(\sum_ic_{i,j}m_i\).
$$
Let $P:M\to M$ be a projection onto a one-dimensional subspace, say ${\mathbb C} m_0$.
By part (a) there is $a\in A$ with $am=Pm$ for every $m\in M$.
We conclude
$$
0\= \sum_j\alpha_j\(P\(\sum_ic_{i,j}m_i\)\)\=\sum_j\lambda_j\alpha_j(m_0),
$$
where $P\(\sum_ic_{i,j}m_i\)\=\lambda_j m_0$.
For $a\in A$ arbitrary, we get
$
0\=\sum_j\lambda_j\alpha_j(am_0).
$
As $a$ runs through $A$, the vector $am_0$ runs through $M$, so
$
\sum_j\lambda_j\alpha_j\= 0.
$
Since the $\alpha_j$ are linearly independent, all $\lambda_j$ are zero, hence all $\sum_ic_{i,j}m_i$ are zero and again by linear independence, we get $c_{i,j}=0$.
This proof also shows that all simple $A$-submodules of $V$ are isomorphic, whence the uniqueness.
\phantom{}\ensuremath{\hfill\square}
We now prove Theorem \ref{4.5.9}.
Let $\eta$ be a $K\times L$-admissible representation of $G\times H$.
Let $E$ be a finite subset of $\hat K$ and $F$ a finite subset of $\hat L$.
Then $V_\eta(E\times F)$ is a finite-dimensional irreducible $C_{E\times F}$-module.
By Lemma \ref{Lem7.5.22} the space $V_\eta(E\times F)$ is an irreducible $C_E\otimes C_F$-module and by Lemma \ref{Lem7.5.23} the module $V_\eta(E\times F)$ is a tensor product of modules, which we write as $V_\pi(E)\otimes V_\tau(F)$.
The uniqueness of the tensor-factors give injective homomorphisms
$\varphi_E^{E'}:V_\pi(E)\to V_\pi(E')$ if $E\subset E'$ and likewise for $F$.
By the uniqueness of the inner products in the Lemma of Schur, we can scale these homomorphisms in a way that they are isometric.
We define
$$
\tilde V_\pi\ \stackrel{\mbox{\rm\tiny def}}{=}\ \lim_\to V_\pi(E).
$$
Then $\tilde V_\pi$ is a pre-Hilbert-space and we write
$V_\pi$ for its completion.
Analogously we construct $V_\tau$.
For every finite subset $E\subset \hat K$ the algebra $C_E$ acts on $\tilde V_\pi$ by continuous operators.
These extend continuously to $V_\pi$ and we get *-representation of the Hecke-algebra ${\cal H}_G$ and likewise for $V_\tau$.
By construction the isotypes of $V_\pi$ are precisely the spaces $V_\pi(E)$ for $E\in I_K$.
The isometrical maps $V_\pi(E)\otimes V_\tau(F)\hookrightarrow V_\eta$ induce an isometric
${\cal H}_G\otimes {\cal H}_H$-homomorphism $\Phi:V_\pi\otimes V_\tau\to V_\eta$, which by irreducibility must be an isomorphism.
It only remains to install a unitary representation of the group $G\times H$ on $V_\pi\otimes V_\tau$, so at first a representation $\pi$ of $G$ on $V_\pi$.
Fix an arbitrary vector $w\in V_\tau$ with $\norm w=1$.
Then the map $V_\pi\to V_\pi\otimes w\tto\Phi V_\eta$ is an isometric embedding of $V_\pi$ into $V_\eta$, which commutes with the ${\cal H}_G$-operation.
The $G$-representation on $V_\eta$ then defines a unitary $G$-representation on $V_\pi$, which induces the ${\cal H}_G$-representation.
We do the same with $H$ and get irreduzible
unitary representations $\pi$ and $\tau$ of $G$ and $H$ such that $\Phi$ is a $G\times H$-isomorphism.
\phantom{}\ensuremath{\hfill\square}
|
\section{Introduction}
\label{Introduction}
The presence of waves and oscillations in the solar corona is a well known feature that has been observed for long time. For an overview of the early observational background see \citet{T1988}. Nowadays, because of the increasing spatial and temporal resolution of the EUV instruments onboard TRACE, SOHO and HINODE spacecraft, accurate observations of oscillations in different coronal structures are accomplished. Many authors have reported observations of transversal coronal loop oscillations from both ground and space-based instruments \citep{AFS1999,NOD1999,APS2002,SAT2002}. When these observations are compared with theoretical models \citep{REB1984,NOD1999,NO2001}, the possibility of inferring some plasma parameters, otherwise difficult to measure, and of improving the existing theoretical models is open; see \citet{BEO2007} for a review. Magnetohydrodynamics (MHD) is the underlying theory of coronal seismology and it is believed that all these observed oscillations and waves can be interpreted theoretically in terms of MHD modes of different coronal plasma structures.
The theoretical study of these oscillations and waves can be done from several points of view. The first approach is to make a normal mode analysis of the linearized MHD equations, which allows to obtain the spatial distribution of the eigenmodes of the structure together with the dispersion relation $\omega(\bf{k})$. Once the elementary building blocks of the MHD normal mode theory are described, the main properties of the resulting MHD waves can be outlined. Many authors have explored the normal modes of coronal structures, beginning with very simple cases such as the straight and infinite cylinder \citep{ER1983}. In the context of curved coronal magnetic structures, \citet{GPH1985,PHG1985,PG1988} investigated the continuous spectrum of ideal MHD. \citet{OBH1993,OHP1996} and \citet{TOB1999} derived the spectrum of modes in potential and nonpotential arcades. More complex configurations, such as sheared magnetic arcades in the zero-$\beta$ plasma limit, have been studied by \citet{AOB2004,AOB2004a}. Other authors have studied eigenmodes in curved configurations with density enhancements that represent coronal loops \citep[e.g.,][]{VDA2004,TOB2006,VFN2006a,VFN2006b,VFN2006c,DAR2006,VVT2009}.
An alternative approach is to obtain the time dependent solution of the MHD equations. Using this method, \citet{CB1995,CB1995a} studied analytically the propagation of fast waves in a two-dimensional coronal arcade for a particular equilibrium, namely one with uniform Alfv\'en speed. \citet{OMB1998} studied the effect of impulsively generated fast waves in the same coronal structure. \citet{DSV2005} studied the properties of Alfv\'en waves in an arcade configuration, including the transition region between the photosphere and the corona. Other studies have analyzed the effect of the loop structure on the properties of fast and slow waves in two-dimensional curved configurations \citep[see, e.g.,][]{MSN2005,BA2005,BVA2006,SSM2006,SMS2007}, see \citet{T2009} for a review.
The main aim of this paper is to analyze the effect of including three-dimensional propagation on the resulting MHD waves as a first step before considering more realistic situations like the one observed by \citet{VNO2004}, where the effect of three-dimensional propagation is clear. In our model there is no density enhancement like that of a loop and the zero-$\beta$ approximation is assumed, so only the fast and Alfv\'en modes are present. We focus our attention on the mixed properties displayed by the generated MHD waves that arise due to the coupling when longitudinal propagation is allowed. The paper is arranged as follows. In \S~\ref{equilibrium_conf} we briefly describe the equilibrium configuration as well as some of the approximations made in this work. In \S~\ref{linear} we present our derivation of the linear ideal MHD wave equations with three-dimensional propagation of perturbations. In \S~\ref{numerical_method_and_test} the numerical code used in our study is described, together with several checks that have been performed by solving problems with known analytical or simple numerical solution. Our main results are shown in \S~\ref{numerical_res}, where the linear wave propagation properties of coupled fast and Alfv\'en waves in a two-dimensional coronal arcade, allowing three-dimensional propagation, are described. Finally, in \S~\ref{conclusions} the conclusions are drawn.
\section{Equilibrium configuration}
\label{equilibrium_conf}
We model a solar coronal arcade by means of a two-dimensional potential configuration contained in the $xz$-plane in a Cartesian system of coordinates \citep[see][]{OBH1993}. For this $y$-invariant configuration the flux function is
\begin{equation}
A(x,z)=B\Lambda_{B}\cos{\left(\frac{x}{\Lambda_{B}}\right)}\exp{\left(-\frac{z}{\Lambda_{B}}\right)},
\label{eq:flux}
\end{equation}
and the magnetic field components are given by
\begin{displaymath}
B_{x}(x,z)=B\cos\left(\frac{x}{\Lambda_{B}}\right)\exp\left({-\frac{z}{\Lambda_{B}}}\right),
\end{displaymath}
\begin{equation}
B_{z}(x,z)=-B\sin\left(\frac{x}{\Lambda_{B}}\right)\exp\left({-\frac{z}{\Lambda_{B}}}\right).
\label{eq:arccomp}
\end{equation}
In these expressions $\Lambda_{B}$ is the magnetic scale height, which is related to the lateral extent of the arcade, $L$, by $\Lambda_{B}=2L/\pi$, and $B$ represents the magnetic field strength at the photospheric level ($z=0$). The overall shape of the arcade is shown in Figure~\ref{fig:arc}.
In this paper gravity is neglected and the $\beta=0$ approximation is used for simplicity. Therefore, the equilibrium density can be chosen arbitrarily. We adopt the following one-dimensional profile
\begin{equation}
\rho_{0}(z)=\rho_{0}\exp\left({-\frac{z}{\Lambda}}\right),
\label{eq:density}
\end{equation}
where $\Lambda$ is the density scale height and $\rho_{0}$ is the density at the base of the corona. As shown by \citet{OBH1993}, the combination of magnetic field components given by Equation~(\ref{eq:arccomp}) with the density profile given by Equation~(\ref{eq:density}) leads to a one-dimensional Alfv\'en speed distribution in the arcade that can be cast as
\begin{equation}
v_{A}(z)=v_{A0}\exp{\left[-(2-\delta)\frac{z}{2\Lambda_{B}}\right]}.
\label{eq:Alfven1}
\end{equation}
Here $\delta=\frac{\Lambda_{B}}{\Lambda}$ represents the ratio of the magnetic scale height to the density scale height and $v_{A0}$ is the Alfv\'en speed at the base of the corona. The $\delta$ parameter completely determines the behavior of the Alfv\'en speed profile and hence the wave propagation properties. The case $\delta=2$ represents a uniform Alfv\'en speed model, while $\delta=0$ corresponds to an exponentially decreasing Alfv\'en speed in a uniform density configuration. Other values of $\delta$ represent situations in which both the Alfv\'en speed and density depend on height in a different manner.
\section{Linear waves}
\label{linear}
In order to study small amplitude oscillations in our potential arcade the previous equilibrium is perturbed. For linear and adiabatic MHD perturbations in the zero-$\beta$ approximation the relevant equations are
\begin{equation}
\rho_{0}\frac{\partial\mathbf{v}_{1}}{\partial t}=\frac{1}{\mu_{0}}(\nabla\times\mathbf{B})\times\mathbf{B}_{1}+\frac{1}{\mu_{0}}(\nabla\times\mathbf{B}_{1})\times\mathbf{B},
\label{eq:momentum}
\end{equation}
\begin{equation}
\frac{\partial \mathbf{B}_{1}}{\partial t} =\nabla\times(\mathbf{v}_{1}\times\mathbf{B}),
\label{eq:induction}
\end{equation}
where $\mu_{0}$ is the magnetic permeability of free space and the subscript ``$1$'' is used to represent perturbed quantities. These equations are next particularized to our two-dimensional potential arcade equilibrium. As the equilibrium is invariant in the $y$-direction, we can Fourier analyze all perturbed quantities in the $y$-direction by making them proportional to $\exp{(ik_{y}y)}$. In this way, three-dimensional propagation is allowed and each Fourier component can be studied separately. As a result of this Fourier analysis the perpendicular perturbed velocity and magnetic field components appear accompanied by the purely imaginary number $i=\sqrt{-1}$. This is undesirable from a practical point of view, since Equations~(\ref{eq:momentum}) and (\ref{eq:induction}) will be solved numerically and the code is designed to handle real quantities only. Nevertheless, by making the appropriate redefinitions, namely $v_{1y}\equiv i\tilde{v}_{1y}$ and $B_{1y}\equiv i\tilde{B}_{1y}$, it turns out that our wave equations can be cast in the following form
\begin{eqnarray}
\frac{\partial v_{1x}}{\partial t}&=&\frac{1}{\mu_{0}\rho_{0}}\Bigg[\left(\frac{\partial B_{1x}}{\partial z}-\frac{\partial B_{1z}}{\partial x}\right)B_{z}\Bigg],\label{eq:velocityx}\\
\frac{\partial \tilde{v}_{1y}}{\partial t}&=&\frac{1}{\mu_{0}\rho_{0}}\Bigg[\left(B_{x}\frac{\partial\tilde{B}_{1y}}{\partial x}+B_{z}\frac{\partial\tilde{B}_{1y}}{\partial z}\right)+k_{y}\left(B_{1x}B_{x}+B_{1z}B_{z}\right)\Bigg],\label{eq:velocityy}\\
\frac{\partial v_{1z}}{\partial t}&=&-\frac{1}{\mu_{0}\rho_{0}}\Bigg[\left(\frac{\partial B_{1x}}{\partial z}-\frac{\partial B_{1z}}{\partial x}\right)B_{x}\Bigg],\label{eq:velocityz}\\
\frac{\partial B_{1x}}{\partial t}&=&-k_{y}\tilde{v}_{1y}B_{x}-\frac{\partial}{\partial z}\left(v_{1z}B_{x}-v_{1x}B_{z}\right), \label{eq:fieldx}\\
\frac{\partial \tilde{B}_{1y}}{\partial t}&=&\frac{\partial}{\partial z}\left(\tilde{v}_{1y}B_{z}\right)+\frac{\partial}{\partial x}\left(\tilde{v}_{1y}B_{x}\right),\label{eq:fieldy}\\
\frac{\partial B_{1z}}{\partial t}&=&-k_{y}\tilde{v}_{1y}B_{z}+\frac{\partial}{\partial x}\left(v_{1z}B_{x}-v_{1x}B_{z}\right).\label{eq:fieldz}\\
\nonumber
\end{eqnarray}
These equations constitute a set of coupled partial differential equations with non-constant coefficients that describe the propagation of fast and Alfv\'en waves. As the plasma $\beta=0$, slow waves are excluded from the analysis. When $k_{y}=0$, equations~(\ref{eq:velocityx})--(\ref{eq:fieldz}) constitute two independent sets of equations. The two equations for $\tilde{v}_{1y}$ and $\tilde{B}_{1y}$ are associated to Alfv\'en wave propagation. On the other hand, the four equations for the remaining variables, $v_{1x}$, $v_{1z}$, $B_{1x}$, $B_{1z}$, describe the fast wave propagation. The basic normal mode properties of fast and Alfv\'en modes in a potential arcade with $k_{y}=0$ are described in \citet{OBH1993}, while the case $k_{y}\neq0$ has been considered by \citet{AOB2004}. The time dependent propagation for $k_{y}=0$ was analyzed by \citet{TOB2008}. When longitudinal propagation of perturbations is allowed ($k_{y}\neq0$), the six equations and their solutions are coupled so we may anticipate fast and Alfv\'en wave propagation to display a mixed nature, in an analogous way to the mixed character of eigenmodes obtained by \citet{AOB2004} in their analysis of the normal modes of the present equilibrium with $k_{y}\neq0$. In the following the tildes in $\tilde{v}_{1y}$ and $\tilde{B}_{1y}$ are dropped.
\section{Numerical method and test cases}
\label{numerical_method_and_test}
\subsection{Numerical method}
\label{numerical_method}
The set of differential equations (\ref{eq:velocityx})--(\ref{eq:fieldz}) is too complicated to have analytical or simple numerical solutions except for simplified configurations and under particular assumptions. For this reason we solve them by using a numerical code, although comparisons with known wave properties have been carried out whenever possible.
When considering a potential arcade as the equilibrium magnetic field, it is advantageous to use field-related components instead of Cartesian components in order to characterize the directions of interest related to the polarization of each wave type. The unit vectors in the directions normal, perpendicular, and parallel to the equilibrium magnetic field are given by
\begin{equation}
\mathbf{\hat{e}}_{n}=\frac{\nabla A}{\mid\nabla A\mid},\quad\mathbf{\hat{e}}_{\bot}=\mathbf{\hat{e}}_{y},\quad\mathbf{\hat{e}}_{\Vert}=\frac{\mathbf{B}}{\mid\mathbf{B}\mid},
\label{eq:genunitvec}
\end{equation}
where $A$ is the flux function given in Equation~(\ref{eq:flux}). These unit vectors are related to the Cartesian ones as follows
\begin{equation}
\mathbf{\hat{e}}_{n}=\frac{(B_{z}\mathbf{\hat{e}}_{x}-B_{x}\mathbf{\hat{e}}_{z})}{\mid\mathbf{B}\mid},\quad\mathbf{\hat{e}}_{\bot}=\mathbf{\hat{e}}_{y},\quad\mathbf{\hat{e}}_{\Vert}=\frac{(B_{x}\mathbf{\hat{e}}_{x}+B_{z}\mathbf{\hat{e}}_{z})}{\mid \mathbf{B}\mid},
\label{eq:potentialunitvec}
\end{equation}
with $\mid \mathbf{B}\mid=(B_{x}^2+B_{z}^2)^{1/2}$. In the absence of longitudinal propagation (i.e. for $k_{y}=0$), these three directions are associated with the three types of waves that can be excited, namely $v_{1n}$ for fast waves, $v_{1\bot}$ for Alfv\'en waves, and $v_{1\Vert}$ for slow waves.
Since we want to model a coronal disturbance with a localized spatial distribution we have considered as the initial condition a two-dimensional Gaussian profile given by
\begin{equation}
v_{1}=v_{s}\exp\left[-\frac{(x-x_{s})^2+(z-z_{s})^2}{a^2}\right],
\label{eq:perturbation}
\end{equation}
where $v_{s}$ is the amplitude of the velocity perturbation, $x_{s}$ and $z_{s}$ are the coordinates of the perturbation's center, and $a$ is the width of the Gaussian profile at half height. In the following we use $v_{1}=v_{1n}$ to excite fast waves and $v_{1}=v_{1y}$ to excite Alfv\'en waves. When $k_{y}=0$ the fast mode produces plasma motions purely normal to the magnetic field, while the Alfv\'en mode is characterized by a purely perpendicular velocity component. When propagation along the $y$-direction is considered, pure fast or Alfv\'en modes do not exist and both produce motions in the normal velocity component as well as in the perpendicular velocity component \citep{AOB2004}. It must be noted that the numerical code solves the time-dependent equations in Cartesian coordinates and so the solution has to be transformed following expressions (\ref{eq:potentialunitvec}) to the field-related coordinates. The same applies to the initial perturbation, which must be transformed into the corresponding Cartesian components.
The numerical code \citep[see][for details about the method]{BBT2009} uses the so-called method of lines for the discretization of the variables and the time and space variables are treated separately. For the temporal part, a fourth-order Runge-Kutta method is used. For the space discretization a finite-difference method with a fourth-order centered stencil is choosen. For a given spatial resolution, the time step is selected so as to satisfy the Courant condition. As for the boundary conditions, as we computed the time evolution of two initial perturbations, two kinds of boundary conditions are used. First, when the initial perturbation in $v_{1n}$ is the fundamental normal mode of the $k_{y}=0$ problem, for the $v_{1n}$ component line-tying conditions are chosen at all boundaries, while for the $v_{1y}$ component, flow-through conditions are selected except at $z=0$ where line-tying condition is used. On the other hand, when an initial perturbation like (\ref{eq:perturbation}) is considered, the large photospheric inertia is accomplished by imposing line-tying boundary conditions at $z=0$. In all other boundaries flow-through conditions are used so that perturbations are free to leave the system. In order to increase numerical stability, fourth-order artificial dissipation terms are included in the numerical scheme. In all the simulations the effects of this artificial dissipation have been checked to ensure that they do not affect the obtained solution, but just contribute to eliminate undesired high-frequency numerical modes.
\subsection{Test cases}
\label{test_cases}
Some preliminary tests have been performed in order to figure out the appropriate values of numerical parameters, such as the grid resolution or the numerical dissipation, on the obtained results for fast and Alfv\'en waves. The first test we have conducted has been to run the code with no perturbation at all and to check that the structure remains stable. The results of this numerical run were completely satisfactory. Then the propagation of linear fast and Alfv\'en MHD waves in a potential coronal arcade has been considered.
\subsubsection{Fast wave}
\label{fast_wave}
The temporal evolution of impulsively generated perturbations with rather similar conditions has been accomplished by several authors: \citet{CB1995} obtained analytical expressions for the temporal evolution of perturbations when a coronal arcade is taken as the equilibrium state; \citet{CB1995a, OMB1998} numerically computed such solutions when different initial perturbations are used; and more recently \citet{TOB2008} showed the main properties of the time evolution of fast and Alfv\'en waves in low-$\beta$ environments. These works facilitate the comparison of our numerical results with known results as well as with analytical ones.
As shown by \citet{TOB2008}, when different resolutions are used time dependent results reveal that the grid resolution in our two-dimensional domain is not a critical factor for the proper computation of fast waves and that a good representation of the temporal evolution of perturbations can be achieved even with a rather modest resolution of $50\times50$ grid points in the $(x,z)$-plane. As mentioned above, numerical dissipation is introduced in our code in order to ensure numerical stability. This dissipation is proportional to an adjustable parameter, or dissipation factor, $\sigma_{n}$. We have conducted numerical simulations for different values of the dissipation factor and it turns out that the temporal evolution of fast wave perturbations is not modified.
\subsubsection{ Alfv\'en wave}
\label{alfven_wave}
The properties of Alfv\'en continuum normal modes in a potential coronal arcade described by \citet{OBH1993} allow us to anticipate and identify possible sources of difficulties in the numerical computation of Alfv\'en wave solutions. First of all, since they are oscillatory solutions strongly confined around given magnetic surfaces (both when propagating or in their standing mode version), spatial scales quickly decrease with time and so we can expect a rather important dependence of the numerical solutions on the number of grid points used to cover the area in and around the excited magnetic surfaces. The situation becomes even worse if we take into account that computations in a Cartesian grid do not allow us to locate all the grid points along magnetic surfaces. This fact affects the numerical results and adds a numerical damping. Furthermore, when time-dependent simulations are considered the sampling rate is no more an independent parameter. When the spatial resolution of the grid is defined, the Courant condition gives a maximum value for the temporal resolution which in turn sets the maximum frequency that can be resolved.
We have first generated Alfv\'en waves in our potential arcade model by considering an impulsive initial excitation of the $v_{1y}$ component given by Equation~(\ref{eq:perturbation}) with $z_{s}=1$ and $x_{s}=0$. This implies that the initial disturbance is even about $x=0$ and so odd Alfv\'en modes are not excited. As described by \citet{TOB2008}, the spatial resolution of the numerical mesh affects the obtained amplitude and frequency values. Better resolution provides a closer value to the analytical frequency and less numerical damping. We have also checked the influence of numerical dissipation and the results show that only the amplitude, and therefore the damping time, decreases when the $\sigma_{n}$ parameter is decreased. The spectral analysis of these oscillations at different heights in the structure is shown in Figure~\ref{fig:spectrumalfvenky0}a. The resulting power spectrum is compared to the Alfv\'en continuum frequencies obtained by \citet{OBH1993}. The frequency associated with the generated Alfv\'en waves coincides with the theoretical normal mode frequencies of the system, which gives us further confidence on the goodness of our code. Alfv\'en waves stay confined to the vertical range of magnetic surfaces that were excited by the initial disturbance, since they cannot propagate energy across magnetic surfaces. The initial perturbation is decomposed by the system in a linear combination of normal modes, but keeping the even parity of the initial disturbance with respect to $x=0$, so energy is only found in the fundamental mode, the second harmonic, etc.
In order to better isolate and show the possible numerical artifacts that the code introduces into the numerical solution we have considered a simpler case, the excitation of a particular Alfv\'en mode around a magnetic surface. According to \citet{OBH1993}, Alfv\'en normal mode solutions can be obtained analytically when $\delta=0$. For this reason we now select $\delta=0$.
The initial excitation could now be given by
\begin{equation}
v_{1y}(x,z)=\hat{v}_{1y}(x)\delta\Big[A(x,z)-A(x=0,z_{m})\Big],
\label{eq:normalmode}
\end{equation}
where $\hat{v}_{1y}$ is the regular part of the solution, $A(x,z)$ is the flux function defined by Equation~(\ref{eq:flux}), and $z_{m}$ gives the maximum height of the magnetic field line in which the normal mode is excited. It is important to note that the regular solution has a well-defined parity with respect to the $x$-direction depending on whether $n$ is chosen even or odd. However, since a delta function is difficult to handle from a numerical point of view, our normal mode-like excitation is performed by an initial perturbation of the form
\begin{equation}
v_{1y}(x,z)=\hat{v}_{1y}(x)\exp{\left[-\frac{A(x,z)-A(x=0,z_{m})}{a^2}\right]}.
\label{eq:normalpert}
\end{equation}
For the regular part, $\hat{v}_{1y}(x)$, the fundamental mode with one maximum along the field lines has been chosen. It should be noted that the width, $a$, of the initial perturbation now causes the excitation of several Alfv\'en modes in a set of neighboring magnetic surfaces. It is important to consider an initial velocity profile which is sufficiently localized in the direction transverse to magnetic surfaces so that only a few of them are excited. As we concentrate on the dynamics of a restricted number of field lines around a magnetic surface the consideration of other models, with different values of $\delta$, would change quantitatively the generated frequencies, but not the overall qualitative conclusions shown here.
Figure \ref{fig:spectrumalfvenky0}b shows the temporal evolution of the excited $v_{1y}$ component at a particular location as a function of time for three different values for the width of the initial disturbance. It is clear that three different solutions are obtained. The two corresponding to the largest widths are rather similar, but the one for the smaller width shows a strong damping. It must be said that the exact solution of this ideal system should display no time damping, hence we assert that this is a numerical effect, that cannot be attributed to a real physical damping mechanism. This undesired effect is less important for larger widths of the initial perturbation since, for a given number of grid points, the initial condition is better resolved spatially.
We next fix the width of the initial disturbance, $a$, and vary the spatial resolution in our domain. Figure \ref{fig:resolution} shows several numerical simulations when an initial normal mode-like excitation (Equation~[\ref{eq:normalpert}]) is made at different heights. It is clear that larger spatial resolution provides the more accurately the undamped oscillatory solution.
Also from this analysis we conclude that the spatial resolution is not a factor that should be taken into account in an isolated manner when considering the numerical description of Alfv\'en waves on given magnetic surfaces. Indeed, and because of the Cartesian distribution of grid points in a system of curved magnetic field lines, low-lying magnetic lines are poorly resolved when compared to high-lying magnetic lines for a given grid resolution. This has implications that are worth to be taken into account as can be seen in Figure~\ref{fig:resolution}. If we compare signals in Figure~\ref{fig:resolution}, we can see that, all parameters being the same, closer results to the analytical solution are obtained for higher magnetic field lines. We can therefore assert that for the numerical simulation of Alfv\'en wave properties the resolution of the grid is an important parameter and that it becomes more critical for low-lying magnetic field lines than for higher ones. It should be noted that the conclusions of these tests can also be applied to the case in which an impulsive excitation is set as the initial perturbation.
\section{Numerical results}
\label{numerical_res}
In this section we present the main results from our numerical investigation. For simplicity, first, the temporal evolution of a normal mode-like fast disturbance is analyzed in order to show how and where resonant absorption, due to three-dimensional propagation of perturbations in a non-uniform medium, takes place. It turns out that previous results obtained for the normal modes of coupled fast and Alfv\'en waves in a potential arcade by \citet{AOB2004} can guide us to understand the time evolution of the system and the energy transfer between resonantly coupled modes. Then, a more complex situation is considered by analyzing the time evolution of the initial perturbation given by Equation~(\ref{eq:perturbation}). It should be noted that our first normal mode time evolution analysis has been proof very useful to further better understand the resulting coupling process between both velocity components when a localized impulsive disturbance is used.
\subsection{Resonant damping of fast MHD normal modes in a potential arcade}
\label{normal_like}
In order to gain some insight into the propagation properties of coupled fast and Alfv\'en waves in our configuration, we first study the time evolution caused by an initial disturbance having the spatial structure of a fast normal mode for $k_{y}=0$ (propagation in the $xz$-plane). As shown by \citet{OBH1993}, pure fast modes in a potential arcade are characterized by a global spatial structure determined by the wavenumbers $k_{x}$ and $k_{z}$, which give rise to smooth distributions with a given number of maxima in the $x$- and $z$-directions. This results in a discrete spectrum of frequencies. The frequencies and spatial structure of the fast modes with $k_{y}\neq0$ were computed by \citet{AOB2004}, who showed that perpendicular propagation produces the coupling of the fast normal modes to Alfv\'en continuum solutions, resulting in modes with mixed properties. We have chosen as initial perturbation the velocity perturbation $v_{1n}$ of the fundamental fast mode for $k_{y}=0$, with one maximum in each direction in the $xz$-plane. When $k_{y}=0$ this produces a standing harmonic oscillation of the system, as in an elastic membrane. When $k_{y}\neq0$, this initial perturbation is not a normal mode of the system, but we expect that, the obtained temporal evolution will not differ very much from the actual normal mode of the coupled solution.
Figure \ref{fig:chap4nomarlkyn0} displays the results of such simulation. The first frame for $v_{1n}$ shows the initial spatial distribution of the perturbation. Initially, $v_{1y}$, the velocity component associated to Alfv\'en waves, is zero. As time evolves, a non zero $v_{1y}$ component appears because of the coupling introduced by the three-dimensional propagation. The panels for $v_{1y}$ in Figure~\ref{fig:chap4nomarlkyn0} show that unlike $v_{1n}$ the excited transversal perturbations are not globally distributed in the potential arcade, but only at preferred locations, around a few magnetic surfaces. When the $v_{1n}$ and $v_{1y}$ signals are measured at one of those locations, $x=0$, $z/L=0.35$, it is seen that the amplitude related to the fast-like perturbation decreases in time, while the amplitude of the Alfv\'en-like component of the perturbation increases in time, see Figure~\ref{fig:compvnvy}. This is an indication of the wave energy transfer due to the resonant coupling of the excited fast normal mode to the Alfv\'enic solution around the excited magnetic surface. For long times a decrease in the amplitude of the velocity component, $v_{1y}$, can be appreciated and is attributed to numerical damping, for the reasons explained in \S~\ref{alfven_wave}.
Further confirmation of the resonant wave energy transfer occurring between the modes can be obtained by computing the time evolution of the total energy density in our system. This total wave energy can be computed as
\begin{eqnarray}
\delta E(\mathbf{r},t)&=&\frac{1}{2}\left[\rho_{0}(v_{1x}^2+v_{1y}^2+v_{1z}^2)+\frac{1}{\mu_{0}}(B_{1x}^2+B_{1y}^2+B_{1z}^2)\right]. \label{eq:totalenergy}\\
\nonumber
\end{eqnarray}
The right-hand side panels in Figure~\ref{fig:chap4nomarlkyn0} show the spatial distribution of this quantity as a function of time. The different frames clearly indicate that, initially, the energy is distributed globally around the center of the system, such as corresponds to the initial perturbation we have used. At later times, this energy is transferred to magnetic surfaces around the particular magnetic field line in the arcade where the signals in Figure~\ref{fig:compvnvy} have been measured. The location of this energy deposition is not an arbitrary one. As previous theoretical works on the resonant energy transfer have shown, \citep[e.g., ][]{W1992,HG1993,RGB1997,AOB2004,RW2009}, global fast modes resonantly couple to localized Alfv\'en continuum modes at the magnetic surfaces where the frequency of the fast mode matches that of the corresponding Alfv\'en mode. In our case the spectral analysis of the wave energy densities associated to the normal and perpendicular components, plotted in Figure \ref{fig:spectrumnormalky1}, allow us to confirm the resonant energy transfer at the location where the fundamental fundamental fast mode frequency crosses the Alfv\'en continuum, that exactly corresponds to the magnetic surface where Alfv\'enic oscillations are excited and energy transfer occurs, see Figure \ref{fig:spectrumnormalky1}b. Although the fast mode frequency crosses other Alfv\'en continua, coupling can only occur if the parity of the fast and Alfv\'en eigenfunctions along the field lines is the same, see further details in \citet{AOB2004}. This prevents the coupling with Alfv\'en continuum modes with two extrema along field lines. Even if the coupling with Alfv\'en modes with three extrema along field lines is allowed, we find no signatures of this resonant coupling in the power spectrum analysis nor the wave energy density evolution.
\subsection{Propagation of coupled fast and Alfv\'en disturbances in a potential coronal arcade}
\label{gaussian_like}
Oscillations in coronal magnetic structures are believed to be generated by nearby disturbances, such as flares or filament eruptions. It is clear that such disturbances are far from being a normal mode of a particular structure as our potential arcade. Therefore, we have next considered the impulsive excitation of perturbations by means of a localized disturbance, which is expected to be a better representation of the real phenomena that often trigger waves and oscillations in the solar corona. In particular a Gaussian velocity perturbation is considered and the response of the system is expected to be different from the one described in \S~\ref{normal_like}, since now the initial perturbation is likely to be decomposed in a linear combination of normal modes with different frequencies that will constitute the resulting propagating wave.
We have produced an impulsive excitation of the $v_{1n}$ velocity component of the form given by Equation~(\ref{eq:perturbation}) and have considered $k_{y}\neq0$. Time evolution of the velocity components and the total energy density are displayed in Figure~\ref{fig:gaussd1kyn0}, which shows that the generated wave has both normal and perpendicular velocity components. Note that $v_{1y}=0$ in the absence of $k_{y}$ \citep[see][]{TOB2008}. It is clear in Figure~\ref{fig:gaussd1kyn0} that the perturbed normal velocity component evolution is similar to the one presented by \citet{TOB2008}, for the decreasing Alfv\'en speed model with constant density and $k_{y}=0$. For the normal velocity component, the shape of the wavefront is not circular, due to the fact that perturbations propagate faster toward the photosphere. For large times the front tends to be planar as the initial curvature of the wave packet is lost. As for the perpendicular velocity perturbation that is excited because of the three-dimensional character of the wave, its spatial distribution is highly anisotropic, with the signal concentrated around many magnetic surfaces. A wavefront with fast-like properties, similar to the one present in $v_{1n}$, can also be seen to propagate upwards producing $v_{1y}$ perturbations until it leaves the system. At the end, a collection of Alfv\'enic oscillations are generated in the arcade. By comparing with the results presented in the previous section we can think about them as being generated by the resonant coupling between the fast-like wavefront and several Alfv\'en continuum solutions, instead of the single resonance case shown in \S~\ref{normal_like}. Once excited, magnetic surfaces remain oscillating with their natural period and for large times they are phase-mixed because of the transverse non-uniformity.
We have next analyzed in a quantitative way the effect of $k_{y}$ on the properties of the generated fast-like wavefront and the induced Alfv\'enic oscillations. Regarding the fast-like wavefront, Figure~\ref{fig:posmaxkyn0} (top-panels) shows different snapshots of the cut along $x=0$ of the $v_{1n}$ component for different values of $k_{y}$. These figures indicate that the larger the value of $k_{y}$ the faster the wavefront propagates. The propagation velocity can be measured by plotting the position of the wavefront maximum as a function of time (see Figure~\ref{fig:posmaxkyn0} bottom-left). The time evolution of the wavefront is followed for $t\geqslant t_{0}$ and the initial position of the maximum is denoted by $z_{0}$. For this relatively simple case, the numerical results can be compared with the analytical formula obtained by the integration of the local Alfv\'en speed profile (see Equation~[$5$] in Oliver et al. 1998). The resulting expression is
\begin{equation}
z(t)=\frac{2\Lambda_{B}}{2-\delta}\log\left[ \pm v_{A0}\frac{2-\delta}{2\Lambda_{B}}(t-t_{0})+\exp\left(\frac{2-\delta}{2\Lambda_{B}}z_{0} \right) \right],
\label{eq:localalfven}
\end{equation}
where the $+$ and $-$ signs correspond to upward and downward propagation, respectively. Figure~\ref{fig:posmaxkyn0} (bottom-panels) shows a perfect correspondence between the numerically measured speed and the analytical expression when different models of the solar atmosphere are considered. The increase of the travel speed of fast-like wavefronts when $k_{y}\neq0$ is an important property to be taken into account in the three-dimensional problem.
For the Alfv\'enic oscillations the power spectrum is analyzed in a cut along $x=0$, which allows us to study the power on different magnetic surfaces. Figure \ref{fig:spectrumd1ky1} shows power at a large number of magnetic surfaces, not just around a selected group of field lines around a given magnetic surface, so a wide range of magnetic surfaces are excited because of the coupling. Also, not just the fundamental mode is excited, but also several higher harmonics. All of them have even parity with respect to $x=0$, such as corresponds to the parity of the $v_{1n}$ perturbation and the parity rule for $k_{y}\neq0$ \citep{AOB2004}. When comparing the power spectrum obtained from the numerical solution with the analytical Alfv\'en continuum frequencies given by \citet{OBH1993} for the case $k_{y}=0$, we see that the signal coincides with the analytical curves for $k_{y}=0$. Such as expected, perpendicular propagation has no effect on the frequencies of Alfv\'en waves generated on different magnetic surfaces in the arcade. This is a known result since $\omega\sim(\mathbf{k}\cdot\mathbf{B})$.
As with the normal mode case, a quantitative analysis of the time-evolution of the wave energy of the system helps to better understand the process of energy conversion between fast and Alfv\'en waves. Figure~\ref{fig:gaussd1kyn0} (right-hand side panels), shows the evolution of the total energy density as a function of time. At early stages this quantity shows a clear signature of a fast-like wavefront propagating through the domain. For long times the energy deposition is spatially distributed on the whole system, not only on a single magnetic surface, as in the previous section. Although a large part of the energy leaves the system in the form of fast-like wavefronts, part of the energy remains trapped in the Alfv\'enic oscillations that are resonantly excited in the arcade. The amount of energy trapped in the system can be calculated by the integration of the total energy density (see Equation~[\ref{eq:totalenergy}]) in the whole domain as a function of time. The result is shown in Figure~\ref{fig:energiesdelta} (solid line). For short times the total energy remains almost constant, but when the fast front reaches the boundaries of the system a strong decrease of this quantity is seen. Resonant wave conversion very quickly produces velocity perturbations in the $y$-direction and the energy associated to these Alfv\'enic components grows up to its maximum value before the fast wavefront leaves the system. At later stages, a fraction of around $5-10 \%$ of the initial total energy is retained in the system and the total energy remains almost constant in the subsequent time evolution. We must note that this energy is trapped even in the absence of any density enhancement or wave cavity.
So far, we have used fixed values for the perpendicular propagation wavenumber, $k_{y}$, and the ratio of magnetic to density scale heights, $\delta$. We have next analyzed the influence of these parameters on the obtained results concerning the energy transfer between fast and Alfv\'en waves. Figure~\ref{fig:energiesky} shows the total energy time evolution for different values of $k_{y}$. Several conclusions can be extracted. First, the amount of energy that is trapped by the system in the form of Alfv\'enic oscillations increases with $k_{y}$ and is above $40\%$ for the largest value of this parameter that we have considered. This can be understood in terms of stronger resonant coupling occurring for larger values of $k_{y}$. The relation between the total energy and the energy associated to the $y$-direction also changes with $k_{y}$, in such a way that, while for relatively small $k_{y}$ almost all the energy of the system is stored in oscillations in the $y$-direction, for larger values of $k_{y}$ there is a difference between the total energy and the Alfv\'enic energy for large times. To understand this we need to mention that for $k_{y}\neq0$ Alfv\'en waves have both perpendicular and normal velocity components and so Alfv\'en wave energy is not only contained in the $y$-direction. Although this effect is less visible in the simulations it can be measured, such as shown in Figure~\ref{fig:energiesky}. Note also that for large times the two energy densities decay. This is due to numerical damping, since when very small scales are created the spatial resolution used is not fine enough to handle the localized Alfv\'enic oscillations that are phase-mixed for large times (see \S~\ref{alfven_wave}).
As for the $\delta$ parameter, it controls our model atmosphere, since it allows us to select different ratios of the magnetic scale height to the density scale height. By repeating the previous numerical experiments for two additional values of this parameter the following results are obtained (see Figure~\ref{fig:posmaxkyn0} bottom-panels). Depending on the value of $\delta$, the Alfv\'en speed profile in the vertical direction has a steeper or flatter profile. This means that the time that a fast-like perturbation needs to reach the boundaries of the system and leave it varies with $\delta$. Therefore, the time at which the sudden decrease of the total energy of the system occurs differs for different values of $\delta$, see Figure~\ref{fig:energiesdelta}. However, the fractional amount of wave energy that is transferred to Alfv\'en waves and is trapped in the system does not depend on the model atmosphere we use. Nevertheless, the rate at which the energy transfer occurs does depend in the model atmosphere, such as can be appreciated from the different slopes of the energy in Figure~\ref{fig:energiesdelta}.
\section{Conclusions}
\label{conclusions}
In this paper we have studied the temporal evolution of coupled fast and Alfv\'en waves in a potential coronal arcade when three-dimensional propagation is allowed. Because of the inclusion of three-dimensional dependence on the perturbed quantities, fast and Alfv\'en waves are coupled and the resulting solutions display a mixed fast/Alfv\'en character. The non-uniform nature of the considered medium produces the coupling to be of resonant nature, in such a way that transfer of energy and wave damping occur in the system.
First, the nature of resonant coupling between a fast normal mode of the system and Alfv\'en continuum modes has been analyzed. It is seen that the fast mode with a global nature resonantly couples to localized Alfv\'en waves around a given magnetic surface in the arcade, thus transferring its energy to the later. The position of the resonant surface perfectly agrees with the resonant frequency condition predicted by several authors in previous studies of this kind, and with the parity rules given by \citet{AOB2004}.
Next, the temporal evolution of a localized impulsive disturbance has been analyzed. The inclusion of perpendicular propagation produces an increase in the wave propagation speed for the fast-like wavefront when compared to the purely poloidal propagation case. As in the previous case, perpendicular propagation induces the excitation of Alfv\'enic oscillations around magnetic surfaces, due to the resonant coupling between fast and Alfv\'en waves. Now these oscillations cover almost the whole domain in the arcade, so that the energy of the initial perturbation is spread into localized Alfv\'enic waves. The frequency of the induced Alfv\'enic oscillations is seen to be independent from the perpendicular wavenumber. As time progresses and the initial wavefront leaves the system part of the energy is stored in these Alfv\'en waves which remain confined around magnetic surfaces. Phase mixing then gives rise to smaller and smaller spatial scales, until the numerical code is unable to properly follow the subsequent time evolution. The energy trapping around magnetic surfaces occurs even in the absence of a density enhancement or a wave cavity structure, and is only due to the non-uniformity of the density profile and the magnetic structuring, which lead to a non-uniform Alfv\'en speed distribution.
Finally, the efficiency of this wave energy transfer between large scale disturbances and small scale oscillations has been studied as a function of the perpendicular wavenumber and for different values of the ratio of the magnetic scale height to the density scale height. It is seen that the first factor strongly affects the amount of energy trapped by Alfv\'en waves. The amount of energy trapped by the arcade increases for increasing value of the perpendicular wavenumber. The particular ratio of magnetic to density scale heights determines how fast the available fast wave energy leaves the system and, therefore, the rate at which energy can be transferred to Alfv\'en waves, but not the final amount of energy stored by the arcade in the form of Alfv\'enic oscillations. These $2.5$D simulations should be extended in the future to more realistic $3$D simulations in order to ascertain the applicability of our conclusions to the real wave dynamics observed in coronal structures.
\acknowledgments
The authors acknowledge the Spanish MCyT for the funding provided under project AYA$2006-07637$ and D. Fanning (http://www.dfanning.com/) for his helpful advices about IDL. S.R. also acknowledges MCyT for a fellowship.
|
\section{Introduction}
Bosons, in contrast to fermions, tend to occupy the same state when given the opportunity. This tendency arises from the fundamental quantum commutation relations of bosonic operators, and is responsible, for example, for the creation of laser light and the Bose-Einstein condensate (BEC). In the case of coherent matter waves, such as those in a BEC, stimulated scattering of bosonic atoms enhances the probability for the number of atoms in an already occupied atomic mode (in particular, the ground state of their external motion) to increase with time, and to grow to macroscopic proportions.
At the level of a few individual photons (as opposed to stimulated emission in lasers, in which the presence of many photons is crucial), the propensity of two photons to bunch together was first observed experimentally by Hong, Ou and Mandel (HOM) \cite{hon87}. Specifically, when two separate photons are incident simultaneously upon a 50/50 beam splitter from opposite sides, the two output modes that emerge from the beam splitter each contain either two or zero photons, but never one photon each.
\begin{figure}[h]
\begin{center}
\includegraphics[width=1.8in]{bsf1x.eps}
\caption{Hong-Ou-Mandel interference at a beam splitter: Two single-photon wave packets ($P_1$ and $P_2$) impinge on a semi-transparent mirror ($M$) and the output wave packets are detected ($D_1$ and $D_2$).}
\end{center}
\end{figure}
Let us illustrate this effect in a simple case, which is illustrated in Fig. 1. (The complete theory is given in subsequent sections.) Two input modes (1 and 2) contain one photon each with the same center frequency, temporal width and polarization, which arrive at the beam splitter at the same time. The initial two-photon state is denoted by
\begin{equation}
|\Psi\rangle_{{\rm in}}=a_{1{\rm in}}^\dagger a_{2{\rm in}}^\dagger|{\rm vac}\rangle, \label{1.1}
\end{equation}
where $a_{1{\rm in}}^\dagger$ and $a_{2{\rm in}}^\dagger$ are creation operators for input modes 1 and 2, respectively. (These modes need not be monochromatic, but can be wave packets \cite{blow90,lou00}.) The relation between the input and output mode operators can be written as
\begin{eqnarray}
a_{1{\rm in}}^\dagger&=&\tau a_{1{\rm out}}^\dagger - \rho a_{2{\rm out}}^\dagger, \nonumber\\
a_{2{\rm in}}^\dagger&=&\rho a_{1{\rm out}}^\dagger + \tau a_{2{\rm out}}^\dagger, \label{1.2}
\end{eqnarray}
where the coefficients $\tau$ (transmissivity) and $\rho$ (reflectivity) were assumed to be non-negative, without loss of generality, and satisfy the auxiliary equation $\tau^2 + \rho^2 = 1$. (Usually, one would give the output operators in terms of the input operators, as in the Heisenberg picture, but for our purposes the inverse transformations are more useful.) The input and output operators satisfy the commutation relations $[a_k,a_k^\dagger] = 1$ for k = 1 or 2. All other commutators are zero. By combining Eqs. (\ref{1.1}) and (\ref{1.2}), and using the fact that the vacuum is invariant under the action of the beam splitter, one finds that the output state is
\begin{eqnarray}
|\Psi\rangle_{{\rm out}}&=&
(\tau a_{1{\rm out}}^\dagger-\rho a_{2{\rm out}}^\dagger)
(\rho a_{1{\rm out}}^\dagger+\tau a_{2{\rm out}}^\dagger)
|{\rm vac}\rangle\nonumber\\
&=&\{\tau\rho[(a_{1{\rm out}}^\dagger)^2-(a_{2{\rm out}}^\dagger)^2]+(\tau^2-\rho^2)a_{1{\rm out}}^\dagger a_{2{\rm out}}^\dagger\}|{\rm vac}\rangle. \label{1.3}
\end{eqnarray}
In the case of exactly 50\% beam splitting, for which $\tau=\rho=1/\sqrt{2}$, Eq. (\ref{1.3}) reduces to
\begin{equation}
|\Psi\rangle_{{\rm out}} =
(|2\rangle|0\rangle-|0\rangle|2\rangle)/\sqrt{2}, \label{1.4}
\end{equation}
where $|n\>$ denotes an $n$-photon Fock state. (All the other modes are still vacuum modes.)
In this case the part of the state with one photon in each output port is zero. The physical interpretation of this cancelation is that there are two paths leading to the same final state component $|1\rangle|1\rangle$: Either both photons transmit through the beam splitter or both photons reflect. The probability amplitudes for these two events have opposite signs [$\tau^2$ and $-\rho^2$ in Eq. (\ref{1.3})], so they cancel. The resulting two-photon bunching is a basic property of nature and is at the heart of quantum information processing schemes using linear optics (beam splitters, polarizers, etc.) \cite{klm}.
Notice that this type of HOM interference does not depend on any relative phase between the input states: Unlike coherent states or classical light waves (whose interference properties at a beam splitter certainly depend on the relative phase of the complex amplitudes), Fock states do not possess any physically relevant phase on which interference could depend.
In the original setting \cite{hon87}, for the HOM effect to work, the photons must be in states that are identical in their polarization and spectral degrees of freedom (but their input states differ in their propagation directions). The question arises: Must two photons (or other bosons) have identical properties before they come together for this type of interference to occur? For example, assuming they have identical polarization states, must two photons initially be in identical energy (spectral) states in order to interfere?
In this paper, we show that the answer is negative: By using an active, frequency-shifting beam splitter, one can observe two-photon interference between photons of different color. The active process can be either reflection by a moving mirror (for small frequency shifts), or three-wave mixing in an optical crystal \cite{lou61,van04,alb04,rou04} or four-wave mixing in an optical fiber (for large frequency shifts) \cite{mck04,mck05,gna06,mec06}. The HOM effect is due to destructive interference in the output state, not the input state. We derive conditions on the forms of the input states that are required for high-visibility two-photon interference. These phenomena, in addition to demonstrating a fundamental property of light, will have applications in quantum information experiments involving photons of different color \cite{rod04,fer04,pfi04,mck08}.
\section{Passive Beam Splitters}
Henceforth, we will use a simplified notation: Instead of writing relations that define the input operators in terms of output operators, as in Eq. (\ref{1.2}), we will simply write $a_1^\dagger \mapsto \tau a_1^\dagger - \rho a_2^\dagger$ etc., to reflect how the input state changes to the output state in the most straightforward way. The operators $a_1$ and $a_2$ appearing on the right-hand side are thus meant to be output operators.
If the beamsplitter is a passive device that preserves the energy and spectrum of each photon separately, then the unitary transformation between the input and output channels can be written as
\begin{eqnarray}
a_1^\dagger(\omega) &\mapsto &\tau a_1^\dagger(\omega) - \rho a_2^\dagger(\omega), \nonumber \\
a_2^\dagger(\omega) &\mapsto &\rho a_1^\dagger(\omega) + \tau a_2^\dagger(\omega). \label{2.1}
\end{eqnarray}
Here we use creation operators $a_1^\dagger(\omega)$ for monochromatic modes with frequency $\omega$ impinging on one input port of the beam splitter and $a_2^\dagger(\omega)$ for modes impinging on the other input port. (We exploit a quasi one-dimensional picture of propagating single-photon wave packets. This picture is valid in the paraxial limit and when the radiation is sufficiently narrowband \cite{blow90,lou00}.) These operators satisfy the commutation relations $[a_k(\omega),a_k^\dagger(\omega')] = \delta(\omega - \omega')$ for $k = 1$ or 2. All other commutators are zero. Suppose now we have an input state containing two photons, one impinging on each input port, described by spectral amplitudes $\phi_1(\omega_1)$
and $\phi_2(\omega_2)$, respectively, which satisfy the normalization conditions
\begin{equation}
\int d\omega_k|\phi_k(\omega_k)|^2=1. \label{2.2}
\end{equation}
Then the input state is
\begin{equation}
|\Psi\>_{{\rm in}} = \int d\omega_1 \phi_1(\omega_1)
a_1^\dagger(\omega_1)
\int d\omega_2
\phi_2(\omega_2)
a_2^\dagger(\omega_2)|{\rm vac}\rangle. \label{2.3}
\end{equation}
This input state is mapped by the beam-splitter transformation onto the output state
\begin{eqnarray}
|\Psi\>_{{\rm out}} = \int\int \hspace{-0.1in}&&d\omega_1 d\omega_2 \phi_1(\omega_1)\phi_2(\omega_2)[\tau\rho a_1^\dagger(\omega_1)a_1^\dagger(\omega_2) - \tau\rho a_2^\dagger(\omega_1)a_2^\dagger(\omega_2) \nonumber \\
&&+\ \tau^2a_1^\dagger(\omega_1)a_2^\dagger(\omega_2) - \rho^2 a_1^\dagger(\omega_2)a_2^\dagger(\omega_1)
]|{\rm vac}\rangle. \label{2.4}
\end{eqnarray}
The last two terms can be made to cancel each other under the right conditions, and that complete destructive interference would correspond to HOM interference. The interference is perfect whenever $\tau^2 = \rho^2$ and
\begin{equation}
\phi_1(\omega_1)\phi_2(\omega_2)=\phi_1(\omega_2)\phi_2(\omega_1) \label{2.5}
\end{equation}
for {\em all} frequencies $\omega_1$ and $\omega_2$. This condition can be rewritten as
\begin{equation}
\phi_1(\omega_1)/\phi_2(\omega_1) = \phi_1(\omega_2)/\phi_2(\omega_2). \label{2.6}
\end{equation}
The left-hand and right-hand sides of Eq. (\ref{2.6}) are functions of different variables, implying that they cannot actually depend on those variables. Hence,
\begin{equation}
\phi_1(\omega_1)=C\phi_2(\omega_2), \label{2.7}
\end{equation}
where $C$ is a constant. Normalization of the quantum states gives $|C|=1$, so $C$ is just the phase factor $\exp(i\theta)$ for some phase $\theta$.\ Equivalently, in the time domain, the Fourier transforms of the spectral density functions must satisfy
\begin{equation}
\tilde{\phi}_1(t)=C\tilde{\phi}_2(t) \label{2.8}
\end{equation}
for destructive interference.
From these conditions one may get the impression that it is important to have indistinguishable photons at the input. But that is a little misleading; what counts is indistinguishability at the output, as we will show. The same output state can be reached by two different paths, and it is the interference between the two paths that matters (as {\em always} in quantum mechanics).
\section{Active Beam Splitters}
\subsection{Moving semi-transparent mirror}
Consider a semi-transparent mirror moving to the left in one dimension. In the mirror frame, two channels with the same carrier frequency (color), and opposite propagation directions (right and left) impinge on the mirror [Fig. 2($a$)]. The input and output channels are related by Eqs. (\ref{2.1}). However, in the laboratory frame [Fig. 2($b$)], light initally propagating to the right will be blue-shifted upon reflection, whereas light initially propagating to the left will be red-shifted upon reflection (and transmitted light does not change frequency).
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=2.4in]{tci_f2a.eps} \hspace{0.4in} \includegraphics[width=2.4in]{tci_f2b.eps}
\caption{Interaction of two waves ($r$ and $l$) at a semi-transparent mirror ($M$). In the mirror frame, the waves have the same frequency and opposite directions of propagation (right and left). In the laboratory frame, the waves have different frequencies and directions.}
\label{pump}
\end{center}
\end{figure}
It is convenient to label the right-propagating channel ``red'' and the left-propagating channel ``blue.'' (Of course, in order to shift actual red light to actual blue light and {\em vice versa}, the speed $v$ would have to be of order $c$.) The moving mirror transforms mode operators in the following way:
\begin{eqnarray}
a^\dagger_R(\omega) &\mapsto &\tau a^\dagger_R(\omega) -
(\rho/\sqrt{\alpha})a^\dagger_B(\omega/\alpha), \nonumber \\
a^\dagger_B(\omega) &\mapsto &(\rho\sqrt{\alpha})a^\dagger_R(\alpha\omega) + \tau a^\dagger_B(\omega). \label{3.1.1}
\end{eqnarray}
The constant $\alpha$ is determined by the Doppler effect, which acts twice:
\begin{equation}
\alpha = (1 - \beta)/(1 + \beta) < 1, \label{3.1.2}
\end{equation}
where the normalized mirror speed is $\beta=v/c$. The factors of $\alpha$ in Eqs. (\ref{3.1.1}), which are absent from Eqs. (\ref{2.1}), are required to conserve the number of photons, because red modes in the frequency interval $d\omega_R$ are coupled to blue modes in the interval $d\omega_B = d\omega_R/\alpha$. Transformation (\ref{3.1.1}) is unitary, preserving the commutation relations for the transformed operators. For example,
\begin{eqnarray}
&&[a_B(\omega),a^\dagger_B(\omega')] \mapsto
[\rho\sqrt{\alpha}a_R(\alpha\omega) + \tau a_B(\omega),\rho\sqrt{\alpha}a^\dagger_R(\alpha\omega') + \tau a^\dagger_B(\omega')] = \delta(\omega-\omega'), \nonumber \\
&&[a_B(\omega),a^\dagger_R(\alpha\omega')] \mapsto
[\rho\sqrt{\alpha}a_R(\alpha\omega) + \tau a_B(\omega),\tau a_R^\dagger(\alpha\omega') - (\rho/\sqrt{\alpha})a_B^\dagger(\omega')] = 0, \label{3.1.3}
\end{eqnarray}
where we used $\delta[\alpha(\omega-\omega')] = \delta(\omega-\omega')/ \alpha.$ Although this transformation preserves the number of photons, it does not preserve their energy, thus earning the right to be called active.
Now suppose we have one photon in each input mode, described by wave packets centered around red and blue frequencies, as the input state:
\begin{equation}
|\Psi\rangle_{{\rm in}} = \int d{\omega_R}\phi_R(\omega_R)a^\dagger_R(\omega_R)
\int d{\omega_B} \phi_B(\omega_B)a^\dagger_B(\omega_B)|{\rm vac}\rangle, \label{3.1.4}
\end{equation}
where the spectral amplitudes satisfy the normalization conditions (\ref{2.2}). This state is transformed by the moving mirror into
\begin{eqnarray}
|\Psi\rangle_{{\rm out}} = \int \int \hspace{-0.1in}&&d{\omega_R}d{\omega_B} \phi_R(\omega_R)\phi_B(\omega_B)[\tau a^\dagger_R(\omega_R) -
(\rho/\sqrt{\alpha})a^\dagger_B(\omega_R/\alpha)] \nonumber \\
&&\times [\rho\sqrt{\alpha}a^\dagger_R(\alpha\omega_B) + \tau a^\dagger_B(\omega_B)] |{\rm vac}\rangle. \label{3.1.5}
\end{eqnarray}
The interesting effect is that there can be destructive interference in the output terms where one red and one blue photon emerge. There are two amplitudes for reaching that final state with opposite signs; the (necessary and sufficient) condition for exact destructive interference for all frequencies $\omega_R$ and $\omega_B$ is
\begin{equation}\label{23}
\tau^2\phi_R(\omega_R)\phi_B(\omega_B)=
\rho^2\phi_R(\alpha\omega_B)\phi_B(\omega_R/\alpha). \label{3.1.6}
\end{equation}
This condition implies that $\tau^2 = \rho^2$ and
\begin{equation}\label{24}
\phi_R(\omega_R)/
\phi_B(\omega_R/\alpha)=\phi_R(\alpha\omega_B)/\phi_B(\omega_B) \label{3.1.7}
\end{equation}
for all frequencies $\omega_R$ and $\omega_B$, which in turn implies that
\begin{eqnarray}
\phi_R(\omega_R) &= &C\phi_B(\omega_R/\alpha), \nonumber \\
\phi_B(\omega_B) &= &C^{-1}\phi_R(\alpha\omega_B), \label{3.1.8}
\end{eqnarray}
with $C$ a constant independent of $\omega_R$ and $\omega_B$. These two conditions are really the same. Normalization gives $|C|^2=1/\alpha$, so we can write $C=\exp(i\theta)/\sqrt{\alpha},$
with $\theta$ some constant phase.
In the time domain this condition gives
\begin{equation}
\tilde{\phi}_B(t):=\int d\omega_B \phi_B(\omega_B)\exp(i\omega_B t) = C^{-1}
\int d(\omega_R/\alpha) \phi_R(\omega_R)\exp(i \omega_R t/\alpha), \label{3.1.9} \end{equation}
which means that
\begin{equation}
\sqrt{\alpha}\exp(i\theta)\tilde{\phi}_B(t)=
\tilde{\phi}_R( t/\alpha). \label{3.1.10}
\end{equation}
If the input wavepackets satisfy these relations, then the output state of the moving mirror will contain either two red photons moving to the right, or two blue photons moving to the left, but never one of each color.
\subsection{Frequency translation in fiber, simple version}
Now consider four-wave mixing in a fiber. Among the many processes that may occur, we are interested in frequency translation (Bragg scattering), in which two strong, classical pump waves ($p$ and $q$) convert ``signal'' ($s$) photons into ``idler'' ($r$) photons, or {\em vice versa} \cite{mck04,mck05,gna06,mec06}. The frequency-matching condition for this process, which is illustrated in Fig. 3, is
\begin{equation}
\omega_q + \omega_r = \omega_s + \omega_p. \label{3.2.1}
\end{equation}
It follows from Eq. (\ref{3.2.1}) that the difference between the signal and idler frequencies, $\omega_s - \omega_r = \Omega$, equals the difference between the pump frequencies, $\omega_q - \omega_p$. It is convenient to refer to the idler as the ``red'' sideband and the signal as the ``blue'' sideband. (Once again, we use these terms just to indicate two different colors.) If the pumps are monochromatic (CW), then each red frequency is coupled to one blue frequency. If the red frequencies all lie within the phase-matched bandwidth of the convertor, the red-to-blue conversion efficiency is frequency independent. If the pump powers and fiber length are chosen judiciously, the probability for frequency translation to occur for a given photon is 50\%, with the remaining 50\% probability assigned to the photon's frequency staying the same.
\begin{figure}[h]
\begin{center}
\includegraphics[width=2.4in]{tci_f3ax.eps} \hspace{0.4in} \includegraphics[width=2.4in]{tci_f3bx.eps}
\caption{Two classical pumps ($p$ and $q$) can drive a four-wave mixing process in which a single signal ($s$) photon is frequency-converted to a single idler ($r$) photon, or {\em vice versa}.}
\label{pump}
\end{center}
\end{figure}
The transformation between the red and blue modes is \cite{mck04,mck05}
\begin{eqnarray}
a_R^\dagger(\omega_R) &\mapsto &\tau a_R^\dagger(\omega_R) - \rho a_B^\dagger(\omega_R+\Omega), \nonumber \\
a_B^\dagger(\omega_B) &\mapsto &\rho a_R^\dagger(\omega_B-\Omega) + \tau a_B^\dagger(\omega_B). \label{3.2.2}
\end{eqnarray}
(An equivalent transformation governs frequency up-conversion in a crystal \cite{lou61}.) This transformation is unitary. By following the steps between Eqs. (\ref{3.1.4}) and (\ref{3.1.6}), one finds that the HOM interference is perfect when
\begin{equation}
\tau^2\phi_R(\omega_R)\phi_B(\omega_B)=
\rho^2\phi_R(\omega_B-\Omega)\phi_B(\omega_R+\Omega). \label{3.2.3}
\end{equation}
Condition (\ref{3.2.3}) implies that $\tau^2 = \rho^2$ and
\begin{eqnarray}
\phi_R(\omega_R) &= &C\phi_B(\omega_R+\Omega), \nonumber \\
\phi_B(\omega_B) &= &C^{-1}\phi_R(\omega_B-\Omega), \label{3.2.4}
\end{eqnarray}
with $C$ a constant, independent of $\omega_R$ and $\omega_B$. Once again, these two conditions are the same. Normalization gives $|C|=1$, and hence one can write $C=\exp(i\theta)$. In the time domain,
\begin{equation}
\exp(i\theta-i\Omega t)\tilde{\phi}_B(t) = \tilde{\phi}_R(t), \label{3.2.5}
\end{equation}
which means that the temporal shapes should be identical, with an overall frequency difference equal to $\Omega$. Equation (\ref{3.2.5}) provides necessary and sufficient conditions for frequency translation to display HOM interference in this simple case: Two input photons, one red and one blue, traversing a nonlinear fiber with the appropriate amount of frequency translation, lead to an output state which contains either two red photons or two blue photons, but never one of each color.
\subsection{Frequency translation in fiber, general version}
If one of the pumps is pulsed, or both pumps are pulsed, each frequency component of the signal is coupled to many frequency components of the idler and {\em vice versa} \cite{mck09}. In either case, the transformation over all frequencies is of the form
\begin{equation}
a^\dagger(\omega)\mapsto \int d\omega' G(\omega,\omega')a^\dagger(\omega'), \label{3.3.1}
\end{equation}
where the kernel $G$ satisfies the unitarity condition
\begin{equation}
\int d\omega' G(\omega,\omega') G^*(\omega'',\omega')=\delta(\omega-\omega''). \label{3.3.2}
\end{equation}
Let us split the frequency interval into two non-overlapping parts, ``red'' and ``blue,'' and let us correspondingly write $a(\omega)=a_R(\omega_R)$ whenever $\omega$ is a ``red'' frequency, and $a(\omega)=a_B(\omega_B)$ whenever $\omega$ is a ``blue'' frequency. Then we can rewrite the kernel in the block-matrix form
\begin{equation}
\left[ \begin{array}{cc} G_{RR}(\omega_R,\omega'_R) & G_{RB}(\omega_R,\omega'_B) \\
G_{BR}(\omega_B,\omega'_R)&
G_{BB}(\omega_B,\omega'_B) \end{array} \right], \label{3.3.3}
\end{equation}
where the constituent (Green) kernels $G_{RR}$, $G_{RB}$, $G_{BR}$ and $G_{BB}$ describe red-red transmission, blue-red conversion, red-blue conversion and blue-blue transmission, respectively. The Green kernels satisfy the unitarity conditions
\begin{eqnarray} \label{3.3.4}
\int d\omega_R' G_{RR}(\omega_R,\omega'_R) G^*_{RR}(\omega_R'',\omega_R')
+
\int d\omega_B' G_{RB}(\omega_R,\omega'_B) G^*_{RB}(\omega_R'',\omega_B')
&=&\delta(\omega_R-\omega_R''), \nonumber \\
\int d\omega_R' G_{BR}(\omega_B,\omega'_R) G^*_{RR}(\omega_R'',\omega_R')
+
\int d\omega_B' G_{BB}(\omega_B,\omega'_B) G^*_{RB}(\omega_R'',\omega_B')
&=&0,
\end{eqnarray}
and similar conditions in which $R \leftrightarrow B$. In terms of these kernels, the transformations between the red and blue modes are
\begin{eqnarray}
a_R^\dagger(\omega_R) &\mapsto &\int d\omega'_R G_{RR}(\omega_R,\omega'_R)a_R^\dagger(\omega'_R) +
\int d\omega'_B G_{RB}(\omega_R,\omega'_B)a_B^\dagger(\omega'_B), \nonumber \\
a_B^\dagger(\omega_B) &\mapsto &\int d\omega'_R G_{BR}(\omega_B,\omega'_R)a_R^\dagger(\omega'_R) +
\int d\omega'_B G_{BB}(\omega_B,\omega'_B)a_B^\dagger(\omega'_B). \label{3.3.5}
\end{eqnarray}
The interference of one red and one blue photon in the final state is perfectly destructive when
\begin{eqnarray}
\int\int \hspace{-0.1in} &&d \omega_R d\omega_B \phi_R(\omega_R) \phi_B(\omega_B) [G_{RR}(\omega_R,\omega_R')
G_{BB}(\omega_B,\omega_B') \nonumber \\
&&+\ G_{RB}(\omega_R,\omega_B') G_{BR}(\omega_B,\omega_R')] = 0 \label{3.3.6}
\end{eqnarray}
for all frequencies $\omega'_R$ and $\omega'_B$. By rearranging the terms in Eq. (\ref{3.3.6}), so that functions of $\omega_R'$ and $\omega_B'$ appear on different sides, one obtains the interference conditions
\begin{eqnarray}
\int d\omega_R \phi_R(\omega_R)G_{RR}(\omega_R,\omega_R') &=
&C\int d\omega_B \phi_B(\omega_B)G_{BR}(\omega_B,\omega_R'), \nonumber
\\
\int d\omega_B \phi_B(\omega_B)G_{BB}(\omega_B,\omega_B') &= &-C^{-1}\int d\omega_R \phi_R(\omega_R)G_{RB}(\omega_R,\omega_B'), \label{3.3.7}
\end{eqnarray}
where $C$ is a constant, independent of frequency. The first condition states that the spectral amplitude of red output photons arising from red input photons be equal (up to a constant) to the spectral amplitude of red output photons arising from blue input photons. The second condition is similar, but for blue output photons.
In general, though, these conditions will not be fulfilled with given pumps in a given fiber. One needs to design the four-wave mixing process in such a way that with 50\% probability a red photon is either converted to a blue photon or remains a red photon with a possibly altered spectral wave function. The opposite transformations must hold for the incoming blue photon. The fiber dispersion, pump pulses, and the input red and blue wave packets must all be designed properly to ensure that the output red and blue wave packets are identical for either input channel. (This is work in progress.)
More precisely, if one uses the unitarity conditions (\ref{3.3.4}) to rewrite and combine Eqs. (\ref{3.3.7}), one obtains the interference conditions
\begin{eqnarray}\label{cond}
\phi_R(\omega_R)&=&C
\int d\omega_B
K(\omega_R,\omega_B)\phi_B(\omega_B), \nonumber \\
\phi_B(\omega_B)&=&C^{-1}
\int d\omega_R
K^*(\omega_R,\omega_B)\phi_R(\omega_R),\label{3.3.8}
\end{eqnarray}
where the HOM kernel $K(\omega_R,\omega_B)$ is
\begin{equation}
K(\omega_R,\omega_B)=\int d\omega'
[G^*_{RR}(\omega_R,\omega')G_{BR}(\omega_B,\omega') - G^*_{RB}(\omega_R,\omega')G_{BB}(\omega_B,\omega')]. \label{3.3.9}
\end{equation}
Recall that every hermitian kernel can be decomposed in terms of its eigenvalues, which are real, and a single set of eigenfunctions, which are orthonormal. In a similar way, every complex kernel has the Schmidt (singular-value) decomposition \cite{eke95,law00}
\begin{equation}
K(\omega_R,\omega_B) =\sum_n \sigma_n R_n(\omega_R) B_n^*(\omega_B), \label{3.3.10}
\end{equation}
where the coefficients (singular values) $\sigma_n$ are real and non-negative, and the mode functions $\{R_n(\omega_R)\}$ and $\{B_n(\omega_B)\}$ are self-orthonormal.
Subroutines are available, which determine the mode functions numerically. By substituting decomposition (\ref{3.3.10}) into Eqs. (\ref{3.3.8}), one finds that the interference conditions can be fulfilled only if there is at least one unit coefficient among the Schmidt coefficients. If there is exactly one unit coefficient, say $\sigma_1=1$, then one must have $\phi_R(\omega_R) = \exp(i\theta) R_1(\omega_R)$ and $\phi_B(\omega_B) = B_1(\omega_B)$ in order for complete interference to occur. If there are multiple unit coefficients, one can take superpositions of the corresponding modes. If the kernel possesses {\em only} unit coefficients, then one can choose any red spectral amplitude $\phi_R(\omega_R)$ and find a corresponding blue spectral amplitude. The simple cases of the preceding subsections are, in fact, such degenerate cases, whenever $\tau^2 = \rho^2$.
In the general case, which corresponds to pulsed pumps, one expects that only certain red and blue wavepackets, of finite bandwidth and duration, lead to perfect HOM interference. In all of these cases, $C$ must again be the phase factor $\exp(i\theta)$. Thus, the fiber and pump requirements for perfect HOM interference may be stated in a different and more succinct way: The HOM kernel must have at least one unit singular~value.
One can analyze the general case in an alternative way, by giving the Schmidt decompositions of the Green kernels directly, rather than the HOM kernel $K$. By using the results of \cite{was}, which are reviewed in the Appendix of this paper, one finds that
\begin{eqnarray}
G_{RR}(\omega_R,\omega_R')
&=&\sum_n \tau_n V_n(\omega_R) v_n^*(\omega'_R), \nonumber \\
G_{RB}(\omega_R,\omega_B')
&=&-\sum_n \rho_n V_n(\omega_R) w_n^*(\omega'_B), \nonumber \\
G_{BR}(\omega_B,\omega_R')
&=&\sum_n \rho_n W_n(\omega_B) v_n^*(\omega'_R), \nonumber \\
G_{BB}(\omega_B,\omega_B')
&=&\sum_n \tau_n W_n(\omega_B) w_n^*(\omega'_B), \label{3.3.13}
\end{eqnarray}
where the non-negative coefficients $\tau_n$ and $\rho_n$ satisfy the auxiliary equations $\tau_n^2 + \rho_n^2 = 1$, and the sets of mode functions $\{V_n\}$, $\{v_n\}$, $\{W_n\}$, and $\{w_n\}$ are self-orthonormal.
By combining Eqs. (\ref{3.3.9}) and (\ref{3.3.13}), one obtains the alternative decomposition
\begin{equation} K(\omega_R,\omega_B) = 2\sum_n \rho_n\tau_n V_n^*(\omega_R) W_n(\omega_B). \label{3.3.14} \end{equation}
It follows from Eqs. (\ref{3.3.10}) and (\ref{3.3.14}) that $\sigma_n = 2\rho_n\tau_n$, $R_n = V_n^*$ and $B_n = W_n^*$. Notice that $\sigma_n \le 1$, so the HOM condition $\sigma_n = 1$ represents a special (ideal) case.
The coefficients $\tau_n$ and $\rho_n$ can be interpreted as generalized beam-splitter coefficients, with $\tau_n$ playing the role of a transmission coefficient (not changing the color) and $\rho_n$ the corresponding reflection coefficient (changing the color), for each mode $n$. Specifically,
the Schmidt-mode expansions
\begin{eqnarray}
\ a_R^\dagger(\omega)|_{\rm in} &= &\sum_n a_n^\dagger V_n(\omega),
\nonumber \\
\ a_B^\dagger(\omega)|_{\rm in} &= &\sum_n b_n^\dagger W_n(\omega),
\nonumber \\
\ a_R^\dagger(\omega)|_{\rm out} &= &\sum_n c_n^\dagger v_n(\omega),
\nonumber \\
\ a_B^\dagger(\omega)|_{\rm out} &= &\sum_n d_n^\dagger w_n(\omega), \label{3.3.15}
\end{eqnarray}
allow one to rewrite the continuous input and output operators for red and blue light in terms of the discrete operators $\{a_n,b_n,c_n,d_n\}$ \cite{blow90}.
In terms of these specially constructed discrete modes, the output operators are related to the input operators in the simple, pair-wise manner
\begin{eqnarray}
a_n^\dagger &= &\tau_nc_n^\dagger - \rho_nd_n^\dagger, \nonumber \\
b_n^\dagger &= &\rho_nc_n^\dagger + \tau_nd_n^\dagger. \label{3.3.16}
\end{eqnarray}
Equations (\ref{3.3.16}) are equivalent to the two-mode equations (\ref{1.2}). As expected, a unit $\sigma_n$ corresponds to a 50/50 beam-splitter transformation with $\tau_n=\rho_n=1/\sqrt{2}$.
Moreover, description (\ref{3.3.13}) allows us to quantify the effect of imperfections in frequency translation. That is, suppose that the fiber process we actually designed does not have any coefficient $\tau_n$ equal to $1/\sqrt{2}$. Then the best we can do is to use the mode that corresponds to the transmission coefficient that is closest to $1/\sqrt{2}$, say $\tau_1\approx 1/\sqrt{2}$. In fact, if we choose in that case $\phi_R(\omega_R)=V_1^*(\omega_R)$ and $\phi_B(\omega_B)=W_1^*(\omega_B)$, then the amplitude of the component of the output state containing one red and one blue photon is
\begin{equation}
\langle {\rm vac}|a_R(\omega_R)a_B(\omega_B)|\Psi\rangle_{\rm out} = (\tau_1^2-\rho_1^2) V_1^*(\omega_R)W_1^*(\omega_B). \label{3.3.17} \end{equation}
This equation generalizes the corresponding term in Eq. (\ref{1.3}).
The deviation from perfect two-photon interference can be quantified as the probability to find one red and one blue photon (at any red-blue frequency pair) in the output, and (\ref{3.3.17}) shows that this probability is $P_{RB}=(\tau_1^2-\rho_1^2)^2$.
\section{Conclusions}
In this paper, we showed that one can observe interference between two photons of different color. These photons arise independently from separate sources before they come together and interfere. The important point is that both photons end up in the same final spectral state (or mode), even though they began in two spectrally distinct states. It is not necessary for them to be identical initially, except in the obvious sense that they are both photons, which are fundamentally identical particles (if one takes a particle view of photons). For example, if one ``red'' and one ``blue'' photon enter the active interferometer, then either two red or two blue photons exit, but never one of each color.
Observing such bosonic bunching of photons, whose colors are different initially, requires reversible and symmetric inelastic scattering processes. Examples of such processes include three-wave mixing (wavelength down-conversion) in a crystal \cite{van04,alb04,rou04} and four-wave mixing (Bragg scattering) in a fiber \cite{gna06,mec06}. Both types of process can shift photon wavelengths by several hundred nanometers, with high photon-conversion efficiencies. Such two-photon interference should find applications in quantum information science, in schemes using entanglement between optical modes of different color.
\begin{acknowledgments}
MGR wishes to dedicate his contribution to this paper to the memory of Krzysztof Wodkiewicz --- a friend for nearly three decades. He will always value the positive influence that Krzysztof had as a friend, teacher and collaborator. This work was supported in part by NSF grant ECCS-0802109.
\end{acknowledgments}
|
\section{Introduction}
Paraxial beams with non-zero orbital angular momentum, which are exact eigenmodes of the paraxial wave equation \cite{Saleh}, have been of considerable interest \cite{Allen,Allen2,Maier}. In addition to spin angular momentum (polarization), paraxial beams carry an intrinsic orbital angular momentum that depends on their spatial structure. While the transfer of spin angular momentum to matter has long been reported for optically anisotropic media \cite{Beth}, the transfer of orbital angular momentum has newly triggered a wide range of applications. A notable example is particle trapping \cite{Simpson}, whereby particles are set into rotation by coupling with the beam's orbital angular momentum and are thus trapped. Such a coupling of the orbital angular momentum with matter is possible in optically inhomogeneous isotropic media \cite{Beijersbergen}. Thus, a simultaneous independent coupling of both spin and orbital angular momentum with matter is to be expected in a medium which is both optically inhomogeneous and anisotropic \cite{Piccirillo,Marrucci}.
In the past decades, the geometric Berry phase \cite{Berry} acquired by an optical beam (as the Pancharatnam phase \cite{Pancharatnam} or a spin redirection phase \cite{Tomita,Chiao}) has attracted extensive attention. In particular, Bliokh et. al. \cite{Bliokh,Bliokh1,Bliokh2} developed a modified geometrical optics approximation for optically weakly inhomogeneous isotropic media, which contained the Rytov-Vladimirskii rotation law \cite{Rytov,Vladimirskii} and the optical Magnus effect \cite{Zeldovich,Zeldovich2} (also called the optical Hall effect) as manifestations of Berry effects. Such geometric Berry effects have been also derived via the semiclassical wave packet approximation \cite{Onoda,Onoda2,Sawada}, which was originally developed for the study of electronic spin transport in solids \cite{Chang,Sundaram,Culcer}.
By studying the propagation of paraxial beams in optically weakly inhomogeneous isotropic media, Bliokh \cite{Bliokh4} demonstrated that the optical Hall effect becomes more pronounced because of the contribution of the orbital angular momentum Hall effect due to the orbit-orbit interaction. Here, we generalize his results for screw dislocated (amorphous) optical media. Mechanical deformation can result in dislocations that change the optical properties of a medium due to the elasto-optic effect. Such dislocations can be induced by the focused beam itself (see e.g. \cite{Kanehira,Brandstetter}). The study of beam transport provides an indirect way to determine the dislocations as well as the elasto-optical properties of the medium. In the present work, we study paraxial beam propagation parallel to the screw axis of a dislocated amorphous medium that is optically weakly inhomogeneous and isotropic. The paraxial beam contains an optical dislocation, namely the optical vortex along its axis, which is a topological object on the wavefront the beam \cite{Berry2}. Regarding the medium as a continuum, we use the differential geometric theory of defects which is isomorphic to three dimensional gravity \cite{Kleinert,Katanaev}. In this geometric approach, the effect of defects on the three dimensional geometry of the medium is incorporated into a metric. The effect of the screw dislocation on the beam's orbital angular momentum is, thus, shown to change the optical vortex strength, rendering vortex annihilation (for left-handed helical modes) or generation of non-integral vortices \cite{Gotte} possible. This is useful in generating fast switchable helical modes for optical information encoding \cite{Marrucci,Marrucci2}. Furthermore, the dislocation is shown to induce a weak \textit{biaxial} anisotropy in the medium due to the elasto-optic effect. This anisotropy, which is felt more strongly by beams that propagate close to the dislocation line, changes the beam's spin angular momentum as well as causing precession of the polarization. Finally, we derive the equations of motion of the beam and demonstrate how beams with different values of spin and/or angular momentum split in the dislocated medium (the optical Hall effect). Because the beam's singular vortex core can be observed with great accuracy, measurement of the splittings is very reliable. Measuring beam splittings for opposite values of the polarization and orbital angular momentum provides an indirect method for determining the Burgers vector as well as the elasto-optic coefficients of the medium. Determination of the former is demonstrated in a simple example.
\section{Effect of screw dislocation on the beam}
Dislocation is a defect caused by the action of internal stress, which changes the physical properties of the medium, in particular, its optical properties. The displacement vector field, $\textbf{u}$, associated with a screw dislocation line oriented along the $z$-axis of the cylindrical coordinates $(\rho,\varphi,z)$ is $\textbf{u}=(0,0,\beta\varphi)$, where $\beta$ is the magnitude of Burgers vector divided by $2\pi$. The corresponding strain tensor field $S_{ij}=\frac{1}{2}(\partial_i u_j+\partial_j u_i)$ is, thus, given by $S_{\varphi z}=S_{z\varphi}=\beta/2\rho$, other components being zero. Regarding the medium as a continuum, we can use the differential geometric theory of defects which is isomorphic to three dimensional gravity \cite{Kleinert,Katanaev}. In this geometric approach, the effect of defects on the three dimensional geometry of the medium is incorporated into a metric. In particular, a screw dislocation, which corresponds to a singular torsion along the defect line, is described by the following metric \cite{Tod}
\begin{equation}
ds^2=d\rho^2+\rho^2 d\varphi^2+(dz+\beta d\varphi)^2. \label{metric}
\end{equation}
This metric carries torsion but no curvature and describes a screw dislocated medium.
We consider a monochromatic paraxial beam with definite values of the spin and orbital angular momentum, propagating in a weakly inhomogeneous isotropic medium where the refractive index $n(\textbf{x})$ varies adiabatically with position ($\nabla n \rightarrow 0$). With the screw axis $z$ as the paraxial direction, the beam's electric field can be represented as in a homogeneous medium according to
\begin{equation}
\textbf{E}_{\sigma lm}(\rho,\varphi,z)=\textbf{e}_\sigma E_{lm}(\rho,\varphi)e^{ik z}. \label{soln}
\end{equation}
Here $k=k_0n$ is the wave number, the adiabatic variation of which has been neglected ($k_0$ being the wave number in vacuum),
$$\textbf{e}_\sigma =\frac{1}{\surd 2}(\textbf{e}_\rho -i\sigma \textbf{e}_\varphi)$$
with $\sigma=\pm1$ ($\textbf{e}_\rho , \textbf{e}_\varphi$ being the cylindrical unit vectors) are the orthonormal polarization vectors corresponding, respectively, to right/left circular polarization and
$$
E_{lm}(\rho,\varphi)=R_{l|m|}(\rho) e^{im\varphi}
$$
$R_{l|m|}$; with $l=0,1,2,\ldots$ and $m=-l,-l+1,\ldots, l$; being the radial solution of the Helmholtz equation ($\sigma$ and $m$ represent the helicity and the $z$-component of the orbital angular momentum of the photon, respectively ($\hbar=1$)). In the presence of screw dislocation, the Laplacian operator $\nabla^2$ appearing in the Helmholtz equation is to be replaced by
$$
\partial_z^2+\frac{1}{\rho}\partial_\rho(\rho\partial_\rho)+\frac{1}{\rho^2}(\partial_\varphi-\beta\partial_z)^2
$$
which is the Laplacian associated with the metric (\ref{metric}) of the dislocated medium. The solution (\ref{soln}), therefore, acquires an additional phase factor due to the coupling of torsion with angular momentum, according to
\begin{equation}
E_{lm}(\rho,\varphi)= R_{l|m|}(\rho) e^{i(m+\beta k)\varphi}.\label{soln2}
\end{equation}
Furthermore, the relative permittivity tensor $n^2 \delta_{ij}$ acquires an anisotropic part $\Delta_{ij}$ due to the strain field of the dislocation, where (see e.g. \cite{Liu})
$$
\Delta_{ij}=-n^4 p_{ijkl}S_{kl}
$$
$p_{ijkl}$ being the elasto-optic coefficients of the medium. For amorphous media, where only two independent elasto-optic coefficients (customarily denoted by $p_{11}$ and $p_{12}$) exist, the screw dislocation yields
$$
\Delta_{\varphi z}=\Delta_{z\varphi}=-\frac{\beta (p_{11}-p_{12})}{\rho}n^4
$$
other components being zero. The principle refractive indices are, therefore,
$$
n_\rho= n,\ \ n_\varphi=n+\frac{\beta (p_{11}-p_{12})}{2\rho}n^3,\ \ n_z=n-\frac{\beta (p_{11}-p_{12})}{2\rho}n^3
$$
to first order in the perturbation, which is generally sufficiently small. Hence, the elasto-optic perturbation induces a weak \textit{biaxial} anisotropy in the medium, which is felt more strongly by beams that propagate close to the dislocation line (small $\rho$). Propagation of the beam through the dislocated medium, thus, introduces the phases $k_0 n_\rho z$ and $k_0 n_\varphi z$ in the linear polarization states represented by $\textbf{e}_\rho$ and $\textbf{e}_\varphi$, respectively. Hence, in (\ref{soln}), the polarization vector should be rewritten as
$$
\textbf{e}_\sigma =\frac{1}{\surd 2}(\textbf{e}_\rho- i \sigma e^{ik_0\Delta n z} \textbf{e}_\varphi)
$$
where
$$\Delta n(\rho)=n_\varphi-n_\rho=\frac{\beta (p_{11}-p_{12})}{2\rho}n^3
$$
is the induced birefringence. Equivalently,
\begin{equation}
\textbf{e}_\sigma =\frac{1}{\surd 2}(\textbf{e}_\rho- i \sigma e^{i\gamma \cot \theta} \textbf{e}_\varphi) \label{e}
\end{equation}
where $\cot \theta=z/\rho$ and $\gamma=\frac{1}{2}\beta k_0 (p_{11}-p_{12})n^3$.
Equations (\ref{soln}) to (\ref{e}) form our expression for the field of a paraxial beam propagating in an optically weakly inhomogeneous and isotropic amorphous medium with screw dislocation. With the beam scalar product defined by \cite{Allen3}
$$
(\textbf{E}_{\sigma lm}|\textbf{E}_{\sigma' l'm'})=\int \int \textbf{E}_{\sigma lm}^\dagger \textbf{E}_{\sigma' l'm'}\; \rho d\rho d\varphi=\delta_{\sigma \sigma'} \delta_{mm'} \int R_{l|m|}^\star R_{l'|m|}\; 2\pi\rho d\rho
$$
these beams form a complete orthonormal set, in terms of which an arbitrary field distribution can be constructed. The paraxial beam contains an optical dislocation, namely the optical vortex of strength $m$ along its axis, which is a topological object on the wavefront the beam \cite{Berry2}. The $z$-component of the beam's orbital angular momentum (per photon) is given by the expectation value
$$
\int\int E_{lm}^\star(-i\partial_\varphi) E_{lm}\;\rho d\rho d\varphi=m+\beta k.
$$
The fact that its value has changed from $m$ to $m+\beta k$ owes itself to the torque exerted by the strain field of the dislocation. The effect of the screw dislocation on the orbital angular momentum, therefore, is to change the optical vortex strength, rendering vortex annihilation (for negative values of $m$, i.e., left handed helical modes) or generation of non-integral vortices \cite{Gotte} possible. This is useful in generating fast switchable helical modes for optical information encoding \cite{Marrucci,Marrucci2}. Moreover, the $z$-component of the beam's spin angular momentum is \cite{Berry3}
$$
\int\int -i \textbf{e}_z\cdot\textbf{E}_{\sigma lm}^\star\times\textbf{E}_{\sigma lm}\;\rho d\rho d\varphi=-i \textbf{e}_z\cdot\textbf{e}_\sigma^\star\times\textbf{e}_\sigma=\sigma \cos(\gamma \cot \theta).
$$
As expected \cite{Berry3}, the spin component varies along the paraxial direction due to the induced elasto-optic birefringence. Therefore, a screw dislocated medium exerts torques that change both the spin and orbital parts of the beam's angular momentum. The spin change, being attributed to the elasto-optic effect, is more pronounced for beams that propagate close to the dislocation line. Note that, the $z$-component of the beam's total angular momentum,
$$
j=\sigma \cos(\gamma \cot \theta)+m+\beta k
$$
reduces to the constant value $\sigma+m$ in the absence of the screw dislocation ($\beta=0$), as it should.
\section{Paraxial beam dynamics in the dislocated medium}
The adiabatic variation of the wave number $k$, caused by the weak inhomogeneity, has negligible dynamical effect and was, therefore, ignored. However, the adiabatic variation of the beam direction (given by its wave vector $\textbf{k}$) plays a geometric role with nontrivial consequences for the beam dynamics. To incorporate this variation, we treat the coordinate frame used in the previous section as a local frame following the beam direction. As usual, the variation gives rise to a parallel transport law in the momentum space, defined by the Berry connection (gauge potential)
$$
\textbf{A}_{\sigma \sigma'}(\textbf{k})=(\textbf{E}_{\sigma lm}|-i\nabla_\textbf{\scriptsize k}| \textbf{E}_{\sigma' lm}).
$$
Using (\ref{soln}) to (\ref{e}), we obtain
$$
\textbf{A}_{\sigma \sigma'}=[j\delta_{\sigma \sigma'}+i\sigma (\delta_{\sigma \sigma'}-1) \sin(\gamma \cot \theta)]\frac{\cot\theta}{k} \textbf{e}_\varphi
$$
or, in matrix notation,
\begin{equation}
\hat{\textbf{A}}=(m+\beta k+ \hat{{\bf \Sigma}} \cdot\textbf{h})\frac{\cot\theta}{k} \textbf{e}_\varphi \label{m}
\end{equation}
where $\hat{{\bf \Sigma}}=(\hat{\sigma}_1,\hat{\sigma}_2,\hat{\sigma}_3)$ is the Pauli matrix vector and
$$
\textbf{h}(\theta)=(0,\sin(\gamma \cot \theta),\cos(\gamma \cot \theta)).
$$
Equation (\ref{m}) generalizes the result of \cite{Bliokh4} for dislocated (amorphous) media. The first two terms are associated with the parallel transport of the beam's transverse structure \cite{Bliokh4}, while the last describes the parallel transport of the polarization vector along the beam. Note, in particular, the appearance of $\hat{\sigma}_2$ which shows the non-Abelian nature of the polarization transport, to be anticipated in view of the anisotropy \cite{Bliokh5}. The Berry curvature (gauge field strength) associated with this connection is ($\hat{\textbf{A}} \times \hat{\textbf{A}}=0$)
$$
\hat{\textbf{B}}=\nabla_\textbf{\scriptsize k} \times \hat{\textbf{A}}=-(m+\beta k+ \hat{{\bf \Sigma}} \cdot \textbf{D})\frac{\textbf{k}}{k^3}
$$
where $\textbf{D}(\theta)= d(\textbf{h}\cos\theta)/d(\cos\theta)$.
In the course of propagation, the paraxial beam evolves according to $\textbf{E}_{\sigma lm}\rightarrow e^{i\hat{\Theta}}\textbf{E}_{\sigma lm}$, where
\begin{equation}
\hat{\Theta}=\int_C \hat{\textbf{A}} \cdot d\textbf{k}= (m+\beta k)\Theta_0+ \hat{{\bf \Sigma}} \cdot \textbf{I} \label{ph}
\end{equation}
is the geometric Berry phase. Here $C$ is the beam trajectory in momentum space, $\Theta_0=\int_C \cos \theta d\varphi$ is the Berry phase accumulated for $\sigma=1$ and $m=0$ in the absence of dislocation and $\textbf{I}=\int_C \textbf{h} \cos\theta d\varphi=(0,\text{Im} \int_C e^{i\gamma\cot\theta}\cos\theta d\varphi,\text{Re} \int_C e^{i\gamma\cot\theta}\cos\theta d\varphi)$. The first term yields a rotation (through angle -$\Theta_0$) of the beam's transverse structure about the direction of propagation, while the second yields a polarization precession about the direction of $\textbf{I}$. Such a spin precession is characteristic of anisotropic media \cite{Bliokh5}. (In the absence of dislocation, the second term in (\ref{ph}) simply yields the phase factor $e^{i\sigma\Theta_0}$ that leads to the well known Rytov rotation.)
In view of the polarization evolution, the Berry curvature for a given beam is, therefore,
$$
\textbf{B}=(e^{i\hat{\Theta}}\textbf{e}_\sigma)^\dagger\hat{\textbf{B}} (e^{i\hat{\Theta}}\textbf{e}_\sigma)=
\textbf{e}_\sigma^\dagger\hat{\textbf{B}}'\textbf{e}_\sigma
$$
where
\begin{eqnarray}
\hat{\textbf{B}}'=e^{-i\hat{\Theta}}\hat{\textbf{B}} e^{i\hat{\Theta}}=-(m+\beta k+\hat{{\bf \Sigma}}' \cdot\textbf{D})\frac{\textbf{k}}{k^3}\nonumber \\
\hat{{\bf \Sigma}}'=\exp(-i\hat{{\bf \Sigma}} \cdot \textbf{I}) \hat{{\bf \Sigma}}\exp(i\hat{{\bf \Sigma}} \cdot \textbf{I}). \ \ \ \ \ \nonumber
\end{eqnarray}
Using the Baker-Hausdorff formula,
$$
e^{-iG}Ae^{iG}=A-i[G,A]+\frac{(-i)^2}{2!}[G,[G,A]]+\frac{(-i)^3}{3!}[G,[G,[G,A]]]+\ldots
$$
and $[\sigma_i ,\sigma_j]=2i\epsilon_{ijk}\sigma_k$, after some calculations we find
$$
\hat{{\bf \Sigma}}'=\hat{{\bf \Sigma}}\cos 2I+\frac{\textbf{I}(\hat{{\bf \Sigma}}\cdot \textbf{I})}{I^2}(1-\cos 2I)+\frac{\hat{{\bf \Sigma}}\times \textbf{I}}{I} \sin 2I
$$
where $I=|\textbf{I}|=|\int_C e^{i\gamma\cot\theta}\cos\theta d\varphi|$.
Hence
$$
\hat{\textbf{B}}'=-(m+\beta k+\hat{{\bf \Sigma}} \cdot\textbf{D}')\frac{\textbf{k}}{k^3}
$$
where
$$
\textbf{D}'=\textbf{D}\cos 2I+\frac{\textbf{I}(\textbf{D}\cdot \textbf{I})}{I^2}(1-\cos 2I)-\frac{\textbf{D}\times \textbf{I}}{I} \sin 2I.
$$
Therefore,
$$
\textbf{B}=-(m+\beta k+\sigma D'_3)\frac{\textbf{k}}{k^3}
$$
with
$$
D_3'=D_3\cos 2I+\frac{I_3(\textbf{D}\cdot \textbf{I})}{I^2}(1-\cos 2I)
$$
which reduces to the result of \cite{Bliokh4} in the absence of dislocation, namely, the field of a magnetic monopole of charge $m+\sigma$ situated at the origin of the momentum space.
The equations of motion of the beam in the presence of momentum space Berry curvature have been derived repeatedly for various particle beams (photons \cite{Bliokh,Bliokh1,Bliokh2,Zeldovich2,Onoda,Onoda2,Sawada}, phonons \cite{Bliokh6,Torabi,Mehrafarin} and electrons \cite{Chang,Sundaram,Culcer,Berard}). We have
\begin{eqnarray}
\dot{\textbf{k}}=k\nabla \ln n \ \ \nonumber\\
\dot{\textbf{r}}=\frac{\textbf{k}}{k}+\textbf{B}\times \dot{\textbf{k}}\nonumber
\end{eqnarray}
where dot denotes derivative with respect to the beam length. These differ from the standard ray equations of the geometrical optics by the term involving the Berry curvature, which yields the beam displacement
\begin{equation}
\delta \textbf{r}=-\int_C (m+\beta k+\sigma D'_3)\frac{\textbf{k}\times d\textbf{k}}{k^3}. \label{def}
\end{equation}
The displacement, which results in the splitting of beams with different polarizations and/or orbital angular momentums, is orthogonal to the beam direction and produces a current across the direction of propagation. This is the optical Hall effect in the dislocated medium and generalizes the result of \cite{Bliokh4}. Because the beam's singular vortex core can be observed with great accuracy, measurement of the displacements is very reliable. Measuring $\delta\textbf{r}$ for opposite values of $m$ and $\sigma$ provides an indirect method for determining the Burgers vector as well as the elasto-optic coefficients (or rather their difference $p_{11}-p_{12}$) of the medium.
\section{Example: circular waveguide}
We conclude by describing a simple situation that demonstrates the application of the result (\ref{def}) in determining the magnitude of the Burgers vector.
Silicon, being transparent in the near infrared, has attracted extensive attention in optoelectronic devices. Waveguides in silicon can serve as optical interconnects in silicon integrated circuits, or distribute optical clock signals in a microprocessor. Fabrication of such waveguides requires a core with a higher refractive index than that of crystalline silicon. Amorphous silicon, with a higher refractive index in the near infrared, thus provides an interesting candidate for the
core material \cite{Dood}. Let us, therefore, consider an amorphous silicon cylindrical waveguide with screw dislocation along its axis $z$. For a beam propagating inside the waveguide along its circular cross section, the trajectory $C$ in (\ref{def}) is a circle of radius $k=2\pi/\lambda$ in the $xy$-plane ($\theta=\pi/2$). Thus, $I=0$ and $D'_3=D_3=1$ so that (\ref{def}) yields, for every one revolution of the beam,
$$
\delta \textbf{r}=-[b+\lambda (m+\sigma)] \textbf{e}_z
$$
where, of course, $\textbf{e}_z$ is the unit vector in the $z$-direction and $b=2\pi \beta$ is the magnitude of the Burgers vector. Therefore, for two beams of opposite polarization and orbital angular momentum, we have a displacement (per unit revolution) in the negative z-direction given by
$$
\delta z^\pm=b\pm\lambda (|m|+1).
$$
While $\delta z^+>0$, $\delta z^-$ can take any value (including zero) by suitably tuning $\lambda$ and $|m|$. Thus one can have the oppositely polarized beams emerging from the same end of the waveguide, with one lagging behind the other, or from opposite ends (see FIG.~\ref{fig1}). Keeping the emerging direction of the right-handed beam as the reference direction, the emerging direction of the left handed beam changes at the threshold value $\lambda (|m|+1)=b$. With $\lambda$ in the near infrared, this provides an experimental method of determining Burgers vector magnitudes of the order of micrometers.
\begin{figure}
\includegraphics{fig1.eps}
\caption{
Propagation of oppositely polarized beams inside a screw dislocated circular waveguide. (i) $\sigma=1, m=|m|$, (ii) $\sigma=-1, m=-|m|$ with $\lambda (|m|+1)<b$ (left) and $\lambda (|m|+1)>b$ (right).
\label{fig1}}
\end{figure}
|
\section{Introduction}\label{sec:intro}
The problem of dark energy has become one of the most important
issues of the modern cosmology since the observations of type Ia
supernovae (SNe Ia) \cite{Riess:1998cb} first indicated that the
universe is undergoing an accelerated expansion at the present stage
(if assuming that the universe is described by the
Friedmann-Lema\^{i}tre-Robertson-Walker (FLRW) model). Many
cosmologists believe that the identity of dark energy is the
cosmological constant which fits the observational data very well.
However, one still has reasons to dislike the cosmological constant
since it suffers from the theoretical problems such as the
``fine-tuning'' and the ``cosmic coincidence'' puzzles \cite{DErev}.
Thus, a variety of proposals for dynamic dark energy have emerged.
For example, the ``scalar field" model \cite{quintenssence} has been
explored for a long time. Besides, the ``holographic dark energy"
models \cite{HDE}, which arise from the holographic principle of
quantum gravity theory, has also attracted much attention.
There are also some other theoretical approaches to explain the
current cosmic acceleration. For example, it is argued that the
acceleration of the universe may signify the breakdown of Einstein's
theory of general relativity \cite{fR}. Another interesting idea
\cite{Void} is based on the assumption that the universe is
described by the spherically symmetric, inhomogeneous
Lema\^{\i}tre-Tolman-Bondi (LTB) metric \cite{LTB1} \cite{LTB2}
\cite{LTB3}. Recently, some authors further proposed a possibility
of mimicking the cosmological constant in an inhomogeneous universe
\cite{Mustapha} \cite{CBKH} \cite{Kolb} \cite{AER}. The idea is
that, since cosmological observations are limited on the light cone,
it is possible to reconstruct an inhomogeneous cosmological model
(indistinguishable from the homogeneous $\Lambda$CDM model) to
explain the cosmic acceleration without a cosmological constant. The
formalism of reconstruction was developed in \cite{Mustapha} and was
applied to the $\Lambda$CDM model in \cite{CBKH}. In \cite{Kolb},
the authors constructed a spherically symmetric, inhomogeneous
cosmological model reproducing the luminosity-distance and the
light-cone mass density of $\Lambda$CDM model up to $z=2$.
In this work, we focus on the reconstructed inhomogeneous
cosmological model and investigate whether it is consistent with
cosmological observations. Since in this model $\Lambda$CDM
observations are exactly reconstructed on the light-cone, we should
seek for some observational test \textit{not limited on the
light-cone} to distinguish it from the standard $\Lambda$CDM model.
Fortunately, we find that the age of the universe is an appropriate
touchstone. The cosmic age, which depends on the evolution of the
universe at a comoving position, contains information not limited on
the light-cone (the age is uniform in the $\Lambda$CDM model but may
be dependent on the position in an inhomogeneous cosmological
model). Thus it may reveal the discrepancies between the
$\Lambda$CDM model and the reconstructed inhomogeneous model. On the
other hand, it is rather convenient to use the cosmic age to test
the validity of a specific cosmological model. To do this one may
just compare the result with the age of some old objects in our
universe \cite{Age}. For example, to be consistent the age of our
universe in our position must not be younger than the age of the
oldest stellar in the Milky Way.
This paper is organized as follows. In Sec. \ref{sec:LTB}, following
the procedure of \cite{Kolb}, we introduce the inhomogeneous LTB
model and explain how to reconstruct $\Lambda$CDM observations in
this model. In Sec. \ref{sec:CosmicAge}, we calculate the cosmic age
at our position and compare the result with the age of some old
objects. We summarize in Sec. \ref{sec:Summary}.
\section{LEMA\^{I}TRE-TOLMAN-BONDI MODELS}\label{sec:LTB}
In this section, following the procedure of \cite{Kolb}, we explain
how to reconstruct an inhomogeneous LTB model having the same
luminosity distance and light-cone mass density of the homogeneous
$\Lambda$CDM model.
The LTB models are spherically symmetric cosmological solutions to
the Einstein equations with a dust stress-energy tensor. The general
metric for the LTB models in a synchronous comoving coordinate takes
the form,
\begin{equation}
ds^2=-dt^2+\frac {R^{\prime 2}(r,t)} {1+\beta(r)} dr^2 +
R^2(r,t)d\Omega^2.
\end{equation}
Following \cite{Mustapha,Kolb}, we use the prime superscript to
denote $\partial/\partial r$, and the overdot to denote
$\partial/\partial t$. Noticing that here $r$ is a dimensionless
coordinate, while $R(r,t)$ has the dimension of length. The
Robertson-Walker metric can be recovered after performing
$R(r,t)\rightarrow a(t)r$ and $\beta(r)\rightarrow -kr^2$. Solving
the Einstein Equations one obtains the generalized ``Friedmann
Equations" for $R(r,t)$ and $\rho(r,t)$,
\begin{equation}
\label{E1}
\dot{R}(r,t)=\sqrt{ \beta(r) + \frac{\alpha(r)}{R(r,t)} },
\end{equation}
\begin{equation}
\label{E2} \kappa \rho(r,t) = \frac {\alpha^{\prime}(r)}
{R^{2}(r,t)R^{\prime}(r,t)}.
\end{equation}
And the photon radial null geodesic equation for $\hat{t}(r)$ is
found directly from the LTB metric (we are only interested in the
past light cone),
\begin{equation}
\frac {d\hat{t}(r)} {dr} = - \frac {R^{\prime}(r,\hat{t}(r))}
{\sqrt{1+\beta(r)}}.
\end{equation}
For convenence we denote quantities on the light cone by a hat. Thus
we have,
\begin{equation}
R(r,\hat{t}(r))\equiv\hat{R}; \ \ \
R^{\prime}(r,\hat{t}(r))\equiv\hat{R^{\prime}}; \ \ \
\rho(r,\hat{t}(r))\equiv\hat{\rho}.
\end{equation}
Then let us focus on the reconstruction procedures. The method was
discussed by Mustapha, Hellaby, and Ellis in 1997 \cite{Mustapha},
and was recently applied to the $\Lambda$CDM model in \cite{CBKH}
\cite{Kolb} (A related formalism was developed by \cite{CR}
\cite{YKN}). Following their procedure, we take advantage of a
coordinate freedom and rescale $r$ so that \emph{on the light cone},
\begin{equation}
\hat{R}^{\prime}=H^{-1}_{0}\sqrt{1+\beta(r)}.
\end{equation}
The corresponding coordinate transformation is
$dr_1=\frac{H_0\partial \hat R/\partial r }{\sqrt{1+\beta(r)}}dr$.
In the following, for simplicity we will use $r$ to denote $r_1$.
The redshift takes the form (see Sec. 2.3 in \cite{Mustapha})
\begin{equation}
\frac{d\hat z(r)}{dr}=(1+z)\frac{ \hat {\dot
{R^\prime}}}{\sqrt{1+\beta(r)}}.
\end{equation}
Following the procedure of \cite{Kolb}, we reconstruct
$\hat{d}_L(z)$ and $\hat{\rho}(z)$ of the $\Lambda$CDM model. To do
this we require that on the light cone,
\begin{equation} \label{reconstruction}
(1+z)^2\hat R(z)=(1+z)\int^z_0 \frac {dz_1}
{H_{\Lambda\textrm{CDM}}(z_1)}, \ \ \
\hat{\rho}(z)dV_{LTB}=\rho_{M,\Lambda\textrm{CDM}}dV_{\Lambda\textrm{CDM}},
\end{equation}
where $\rho_{M,\Lambda\textrm{CDM}}(z)$ stands for mass density in
the $\Lambda$CDM model, and $H_{\Lambda\textrm{CDM}}(z)$ stands for
the Hubble constant in the $\Lambda$CDM model. These quantities take
the forms,
\begin{equation} \label{LCDMValue}
\rho_{\Lambda\textrm{CDM}}(z)=3\Omega_mM_p^2H_{\Lambda\textrm{CDM}}(z)^2,\
\
H_{\Lambda\textrm{CDM}}=H_0\sqrt{\Omega_m(1+z)^3+\Omega_{\Lambda}}.
\end{equation}
For simplicity we only consider a flat $\Lambda$CDM model. We use
the notations,
\begin{equation}
\Omega_{m}=\frac{\rho_{m}(0)}{\rho_{C}(0)},
\ \ \ \Omega_{\Lambda}=\frac{\rho_{\Lambda}(0)}{\rho_{C}(0)},\ \ \ \rho_C=3M_p^2H_0^2,\ \ \ \Omega_m+\Omega_{\Lambda}=1 .
\end{equation}
Then from Eq. (\ref{reconstruction}) it is straightforward to derive
the following expressions,
\begin{equation} \label{BasicEq.1}
\hat{R}(z)= \frac {1}{1+z} \int^z_0 \frac
{dz_1}{H_{\Lambda\textrm{CDM}}(z_1)},
\end{equation}
\begin{equation} \label{BasicEq.2}
H^{-1}_0\hat{R}^2(z)\kappa\hat{\rho}(z) \frac{dr}{dz} =
\frac{3\Omega_{m}H^2_0}{H_{\Lambda
\textrm{CDM}}(z)}\big[\int^z_0\frac{dz_1}{H_{\Lambda
\textrm{CDM}}(z_1)}\big]^2 .
\end{equation}
Furthermore, the following three equations can be obtained by
solving the corresponding $z$, $\alpha(r)$ and $\beta(r)$ (Eqs.
(19-21) in \cite{Kolb})
\begin{equation}\label{BasicEq.3}
\frac {dz}{dr} =(1+z) \frac {H_{\Lambda \textrm{CDM}}(z)}{H_0} ,
\end{equation}
\begin{equation}\label{BasicEq.4}
\frac {d\alpha}{dr}=\frac
{1}{2}H^{-1}_0\hat{R}^2(z)\kappa\hat{\rho}(z)\big[\frac{1}{H_0d\hat{R}/dr}\big(1-\frac
{\alpha}{\hat{R}}\big) + H_0 \frac {d\hat{R}}{dr}\big],
\end{equation}
\begin{equation}\label{BasicEq.5}
\beta(r)=(\frac{d\alpha}{dr}\frac{1}{H^{-1}_0\hat{R}^2\kappa\hat{\rho}})^2-1.
\end{equation}
It should be emphasized that for these values of $\alpha(z)$ and
$\beta(z)$, the corresponding LTB model \emph{exactly reproduces}
the luminosity-distance relation and light-cone mass density of the
$\Lambda$CDM model. For convenience, in the following context we
will use ``LTB-$\Lambda$CDM model" to denote this reconstructed LTB
model.
It should be stressed that since the LTB-$\Lambda$CDM model is
reconstructed by mimicking $\Lambda$CDM model, it has the same
luminosity-distance-redshift relation and light-cone
mass-density-redshift relation as the homogeneous $\Lambda$CDM
model. Thus when estimating some physical quantities of the
LTB-$\Lambda$CDM model from cosmological measurements one can just
use the obtained values of $\Omega_{m}$ and $H_{0}$ from the fit of
the $\Lambda$CDM model (i.e. it is not necessary to impose a special
constraint on the LTB-$\Lambda$CDM model). For example, in
\cite{Kolb} the authors just take $\Omega_m$=0.3 to mimicking a flat
$\Lambda$CDM with mass ratio $\rho_{m0}/\rho_{c0}=0.3$.
\section{Cosmic age at our position}\label{sec:CosmicAge}
In the first subsection, we derive the expression of the cosmic age
in the LTB-$\Lambda$CDM model. Then in the second subsection, we
estimate the age from cosmological observations, and compare it with
that in the homogeneous $\Lambda$CDM model. Finally, in the third
subsection, we discuss the validity of the LTB-$\Lambda$CDM model by
considering the astronomical measurements of the age of some old
objects in our universe.
\subsection{Cosmic Age in the LTB-$\Lambda$CDM}
First of all, we investigate the properties of
Eqs.(\ref{E1}),(\ref{E2}),(\ref{BasicEq.1}-\ref{BasicEq.5}) at $r=0$
(corresponds to our position in the universe). Since the function
$R(r,t)$ represents the diameter distance, one expects
$R(r,t)\rightarrow0$ when $r$ approaches zero. So $R(r,t)$ can be
expanded near $r=0$ as,
\begin{equation}
R(r,t)=R_1(t)r+\ldots
\end{equation}
Then from Eq. (\ref{BasicEq.4}) we obtain
\begin{equation} \label{R1}
\alpha(r)=\alpha_3r^3+....,\ \
\beta(r)=\beta_2r^2+...,\ \
\dot{R}_1(t)=\sqrt{\beta_2+\frac{\alpha_3}{R_1(t)}}.
\end{equation}
From Eqs. (\ref{LCDMValue})(\ref{BasicEq.2})(\ref{BasicEq.4}), we
expand $H_{\Lambda \textrm{CDM}}(z),\ \hat r(z),\ \hat R(z)$ to the
leading order and substitute them into Eq. (\ref{BasicEq.4}) to
obtain $\alpha_3$
\begin{equation}\label{alpha3}
H_{\Lambda \textrm{CDM}}(z)=H_0+...,\ \ r=z+...,\ \ \hat{R}(z)=\frac{z}{H_0}+...,\
\ \alpha_3=\Omega_m H^{-1}_0.
\end{equation}
The calculation of $\beta_2$ is also straightforward. From
Eqs.(\ref{BasicEq.4})(\ref{BasicEq.5}) it follows that
\begin{equation} \label{beta2}
\beta(r)=\big( \frac{1}{2} \big[ \frac{1}{H_0d\hat{R}/dr}\big(
1-\frac{\alpha}{\hat R}\big) +H_0\frac {d\hat R}{dr}{} \big]
\big)^2-1.
\end{equation}
Substituting
$\hat{R}^{\prime}(z)=H^{-1}_0+\hat{R}^{\prime}_1z+\hat{R}^{\prime}_2z^2+...$
into Eq. (\ref{beta2}), we obtain
\begin{equation} \label{beta2two}
\beta(z)=(H^2_0\hat{R}^{\prime 2}_1 -
\Omega_m)z^2+...=(H^2_0\hat{R}^{\prime 2}_1 - \Omega_m)r^2+...
\end{equation}
Notice that $\hat R^{\prime}_2$ does not appear in the expression,
so we just have to expand $\hat R^{\prime}$ to the first order.
Using Eqs.(\ref{BasicEq.1})(\ref{BasicEq.3}), it follows that,
\begin{equation}
\frac{d\hat{R}}{dr} = \frac{d\hat{R}}{dz} \frac{dz}{dr} = -
\frac{1}{1+z} \frac{H_{\Lambda \textrm{CDM}}(z)}{H_0} \int^z_0
\frac{dz_1}{H_{\Lambda \textrm{CDM}}(z_1)}+\frac{1}{H_0} = H^{-1}_0
- H^{-1}_0 z+...,\ \ \Rightarrow\ \ \ \ \ \ \hat
R^{\prime}_1=H^{-1}_0.
\end{equation}
Combining with Eq. (\ref{beta2two}) we obtain
\begin{equation} \label{BasicValue2}
\beta_2=1-\Omega_{m}.
\end{equation}
Next we calculate the age of the universe ($t_0-t_{BB}$) at $r=0$.
From Eq. (\ref{R1}) it follows that,
\begin{equation} \label{CosmicAge1}
t_0-t_{BB}=\int^{t_0}_{t_{BB}}dt=\int^{R_1(t_0)}_{R_1(t_{BB})}
\frac{dR_1(t)}{\dot{R}_1(t)},
\end{equation}
where the upper and lower bounds of the integral can be determined
by Eq. (\ref{E2}),
\begin{equation}
R_1(t)=\big( \frac{3\alpha_3}{\kappa \rho(0,t)} \big)^{\frac{1}{3}}.
\end{equation}
Finally, combining the above with
Eqs.(\ref{LCDMValue})(\ref{alpha3}), we obtain the age of the
universe at $r=0$ in the LTB-$\Lambda$CDM model,
\begin{equation}\label{CosmicAge2}
t_0-t_{BB}=\int^{H^{-1}_0}_{0} \frac{dR_1}{\sqrt{
\Omega_{\Lambda}+\Omega_{m}H^{-1}_0/R_1 }}.
\end{equation}
\subsection{Estimate $t_{LTB}$ from Cosmological Observations}
Next we estimate the cosmic age in the LTB-$\Lambda$CDM model and
compare the result with that of the $\Lambda$CDM model. These two
models are both determined by two parameters, the present matter
ratio $\Omega_m$, and the Hubble constant $H_0$. For convenience,
let us denote the cosmic age in the LTB-$\Lambda$CDM model and
$\Lambda$CDM model at $r=0$ as $t_{LTB}(\Omega_m,H_0)$ and
$t_{\Lambda \textrm{CDM}}(\Omega_m,H_0)$. We have [we perform a
parameter transformation $r=H^{-1}_0/(1+z)$ in $t_{\Lambda
\textrm{CDM}}$]
\begin{eqnarray}
t_{LTB}(\Omega_m,H_0)=\int^{H^{-1}_0}_{0} \frac{dR_1}{\sqrt{
\Omega_{m}/(H_0R_1)+1-\Omega_{m} }}, \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \\
t_{\Lambda \textrm{CDM}}(\Omega_m,H_0)=\int^{+\infty}_0
\frac{dz}{H_{\Lambda}(z)(1+z)}=\int^{H^{-1}_0}_0
\frac{dr}{\sqrt{\Omega_m/(H_0r)+(1-\Omega_m)(H_0r)^2}}.
\end{eqnarray}
The only difference between $t_{\Lambda \textrm{CDM}}$ and $t_{LTB}$
lies in the second term in the square root. Since $H_0r<1$ (the
upper bound of the integral is $H^{-1}_0$), it is clearly that
$t_{LTB}$ is smaller than $t_{\Lambda \textrm{CDM}}$ given the same
set of parameters.
Notice that $t_{LTB}(\Omega_m,H_0)$ is proportional to $1/H_0$. To
prove this one can perform a coordinate transformation
$\tilde{R}_1=H_0R_1$, which yields that,
\begin{equation}
t_{LTB}(\Omega_m,H_0)=\frac{1}{H_0}\int^1_0
\frac{d\tilde{R}_1}{\sqrt{ \Omega_{m}/\tilde{R}_1+1-\Omega_{m} }}.
\end{equation}
So we have,
\begin{eqnarray}
\frac{\partial}{\partial H_0} t_{LTB}(\Omega_m,H_0) &=&
-\frac{1}{H_0}t_{LTB}(\Omega_m,H_0)\ <\ 0,\ \ \ \ \\
\frac{\partial}{\partial \Omega_m} t_{LTB}(\Omega_m,H_0) &=&
-\frac{1}{2}\int^{H^{-1}_0}_0 \big(\
\frac{\Omega_m}{H_0r}+1-\Omega_m
\big)^{-3/2}\big(\frac{1}{H_0r}-1\big)\ dr <\ 0,
\end{eqnarray}
Clearly, $t_{LTB}(\Omega_m,H_0)$ has larger values with smaller
values of $\Omega_m$ and $H_0$ parameters. One can easily verify
that the conclusion is the same for $t_{\Lambda
\textrm{CDM}}(\Omega_m,H_0)$.
Now we are ready to calculate the specific values of $t_{LTB}$ with
given values of $H_0$ and $\Omega_m$. Here we refer to the result of
the seven-year Wilkinson microwave anisotropy probe (WMAP)
observations \cite{WMAP7}. From the constraint from
``WMAP7+BAO+$H_0$" \cite{WMAP7}\cite{Riess:H}\cite{BAO2} the WMAP
collaboration provides the best-fit values of $\Omega_{\Lambda}$ and
$H_0$ together with their 1$\sigma$ uncertainties,
\begin{equation}
\Omega_{\Lambda}=0.728^{+0.015}_{-0.016},\ \ H_0=70.4^{+1.3}_{-1.4}
\ \textrm{km/s/Mpc}.
\end{equation}
From their result we can put a constraint on $t_{LTB}$ and
$t_{\Lambda \textrm{CDM}}$ at 1$\sigma$ confidence level (CL). The
result is,
\begin{equation}
t_{LTB}=11.4 \pm 0.3\ \textrm{Gyr},\ \ \ t_{\Lambda
\textrm{CDM}}=13.8 \pm 0.5\ \textrm{Gyr}.
\end{equation}
It is found that $t_{LTB}$ is about 2Gyr younger than $t_{\Lambda
\textrm{CDM}}$. At the 1$\sigma$ CL we obtain the upper limit of the
age of the LTB-$\Lambda$CDM model $t_{LTB}<11.7$Gyr, with the set of
the smallest values of parameters $\Omega_m=0.257$ and
$H_0=69.0$km/s/Mpc (In fact this result is overestimated since we
ignore the degeneracy between $\Omega_m$ and $H_0$. The upper limit
value of $t_{LTB}$ should be smaller, or at least as small as
11.7Gyr).
It should be stressed that to be strict the previous estimation of
$t_{LTB} $ is not appropriate, since we have assumed that the WMAP
and baryon acoustic oscillations data could be used to constrain the
LTB models. In fact, the issues of CMB and structure formation in
the LTB scenario are rather complicated and have not been clearly
investigated. To avoid this problem one can put constraints to
$\Omega_m$ and $H_0$ from SNIa observations. The recent observations
of SDSS-II (Sloan Digital Sky Survey II)\cite{SDSS2} and Hubble
Space Telescope\cite{Riess:H} show that $\Omega_m>0.224,\
H_0>70.6$km/s/Mpc in 1$\sigma$ CL. From their result we find a upper
limit $t_{LTB}<11.6$Gyr, which is the similar with our previous
result $t_{LTB}<11.7$Gyr.
\subsection{Discussions of the Validity of the LTB-$\Lambda$CDM Model}
The low limit to the cosmic age can be directly obtained from
estimating the age of some old objects in our universe
\cite{Richer}\cite{Hansen}\cite{Krauss}. As an example, based on
white dwarf cooling the authors of \cite{Hansen} get a result of
$12.7 \pm 0.7$Gyr. Compared with the result of the previous
subsection $t_{LTB}<11.7$Gyr it seems that the LTB-$\Lambda$CDM
model is inconsistent with their measurements. However, one should
not conclude hastily. The reason is that the result of \cite{Hansen}
is subject to larger uncertainty. The uncertainty due to
calculations of the white dwarf cooling is difficult to estimate,
and in their result corresponding errors are not included. In fact,
the authors of \cite{Hansen} argued that systematic uncertainties
are likely to be at least as large as, if not larger than, the
quoted statistical errors. So if indeed the uncertainty due to
calculations of the white dwarf cooling is as large as the
observational error, then the LTB-$\Lambda$CDM model is within
1$\sigma$ agreement with observations.
The age based on evolution of compact binaries is somewhat lower. In
\cite{Kaluzny}\cite{Chaboyer} the authors give a result of $11.8 \pm
0.6$Gyr and $11.10 \pm 0.67$Gyr, respectively. Moreover, the oldest
known star in the Milky Way, HE 1523-0901, is reported to have an
age of 13.2$\pm$2.7Gyr \cite{HE1523}, for a lower limit of 10.5Gyr.
Obviously, these measurements are all in consistent with $t_{LTB}$
and $t_{\Lambda \textrm{CDM}}$ obtained with data from \cite{WMAP7}
at 1$\sigma$ CL. Therefore, although the obtained $t_{LTB}$ is about
2Gyr younger than $t_{\Lambda \textrm{CDM}}$, it is still in
1$\sigma$ agreement with current astronomical observations, and we
are not able to argue against the reconstructed LTB-$\Lambda$CDM
model.
We show the situation in Fig. 1. The green and red regions represent
parameters with cosmic age older than 11.2Gyr in $\Lambda$CDM and
LTB-$\Lambda$CDM model, respectively. The black shadow region is a
1$\sigma$ CL constraint from the seven-year WMAP observations
\cite{WMAP7} (we ignore degeneracy). The blue region is a 2$\sigma$
CL constraint from a joint analysis performed in one of our previous
works \cite{Constraints}, in which we used the Constitution
supernovae sample \cite{Constitution}, the baryon acoustic
oscillations \cite{BAO} and the five-year WMAP observations
\cite{WMAP5}. Since current limit to the cosmic age from
astronomical measurements generally gives a result $t<11.2$Gyr, we
plot the regions of $t_{LTB}<11.2$Gyr (red shadow) and
$t_{\Lambda\textrm{CDM}}<11.2$Gyr (green shadow) in this figure. It
is obvious that the $\Lambda$CDM perfectly passes the cosmic age
text, while the overlap of the blue region, the black shadow region
and the red shadow region implies that the LTB-$\Lambda$CDM model is
also consistent with current observations.
Finally, at the end of this section, we stress that in this paper we
only consider a particular LTB model - namely, the one with
$\Lambda$CDM features. Other inhomogeneous models (such as the void
models) may have larger age at the origin, and in these cases one
would have to seek for other methods to test and identify them.
\begin{figure}\label{Fig1}
\includegraphics[width=12.3cm]{CosmicAge.eps}
\caption{Parameter space in $\Omega_m$ and $H_0$ plane. The green
and red regions stand for $t_{\Lambda \textrm{CDM}}>$11.2Gyr and
$t_{LTB}>$ 11.2Gyr respectively. The black shadow region represents
the 1$\sigma$ constraint to $\Omega_m$ and $H_0$ from the seven-year
WMAP observations (degeneracy is ignored). The blue region is a
2$\sigma$ constraint from a joint analysis from the Constitution
supernovae sample, baryon acoustic oscillations and the five-year
WMAP observations. }
\end{figure}
\section{Summary}\label{sec:Summary}
In this paper we calculate the cosmic age at $r=0$ in the
LTB-$\Lambda$CDM model, which reproduces the luminosity-distance and
light-cone matter density of the homogeneous $\Lambda$CDM model.
Using the constraints of $\Omega_m$ and $H_0$ from the seven-year
WMAP observations combined with other cosmological observations, we
get the upper limit $t_{LTB}<11.7$Gyr at 1$\sigma$ CL. This result
is about 2Gyr younger than the cosmic age in $\Lambda$CDM scenario.
Since current astronomical measurements generally put a 1$\sigma$ CL
lower limit on the age of the universe of about 11.2Gyr, the
LTB-$\Lambda$CDM model is still in 1$\sigma$ agreement with all
these observations. However, due to the relatively younger age the
LTB-$\Lambda$CDM model might be disfavored by future observations.
Besides, even if the LTB-$\Lambda$CDM model successfully passes all
the tests of future observations, there might be some other problems
in this scenario.(The discussions of these complicated problems are
beyond the scope of this paper.) The main reason is that it is
difficult to fit this model into the larger framework of fundamental
physics such as particle physics, general relativity, astrophysics,
and cosmology. For example, in \cite{Kolb} the authors mentioned the
theory of structure formation and the integrated Sachs-Wolfe effect:
study of these issues is very difficult in the LTB models.
The topic of distinguishing the homogeneous $\Lambda$CDM model and
the reconstructed inhomogeneous LTB-$\Lambda$CDM model is
scientifically interesting and important, since it involves the
question of the mysterious feature of dark energy and whether the
nearby region of the universe is homogeneous. This topic should be
carefully investigated. In this paper we propose the possibility of
distinguishing the LTB-$\Lambda$CDM scenario with the standard
$\Lambda$CDM model by performing the cosmic age test. The procedure
is convenient and straightforward. Although the result shows that
the LTB-$\Lambda$CDM is still in 1$\sigma$ agreement with current
astronomical observations, since with the same set of parameters
this model always has a younger age than the standard $\Lambda$CDM
model it is possible to distinguish them from future observations.
In all, the issue of using the cosmic age test to distinguish the
reconstructed inhomogeneous LTB models from the homogeneous
$\Lambda$CDM model is worth further investigation, and should be
taken into consideration in future works, e.g., in the cases when
people try to construct a new model in the LTB scenario to explain
the apparent cosmic acceleration.
\begin{acknowledgments}
The authors would like to thank the anonymous referees for carefully
examining our paper and providing us a number of important comments.
This work was supported by the Natural Science Foundation of China
under Grants No. 10535060/A050207, No. 10975172 and No. 10821504,
and Ministry of Science and Technology 973 program under Grant No.
2007CB815401. SW also thanks the support from a graduate fund of the
University of Science and Technology of China.
\end{acknowledgments}
|
\section{Introduction}
The influence of high porosity aerogel as quenched disorder has been studied in various systems such as liquid $^4$He,\cite{chan} $^3$He-$^4$He mixture,\cite{kim,McRae} $^3$He,\cite{porto,sprague} and liquid crystals.\cite{bellini} The effect of aerogel on superfluid $^3$He is exceptionally interesting because it is a {\it p-wave triplet} anisotropic superfluid possessing continuous symmetry. Since the discovery of superfluiditiy of $^3$He in high porosity aerogel,\cite{porto,sprague} more than a decade of theoretical and experimental efforts have been invested to understand this system and have revealed many interesting phenomena. The fragile nature of {\it p-wave} pairing against impurity-scattering was immediately recognized by the significant depression of superfluid transition, \cite{porto,sprague,matsumoto} and the theoretical descriptions based on various isotropic impurity-scattering models have provided a successful account for the observed behavior. \cite{thunne,hanninen,sauls03} A wide variety of experimental evidence reflecting the role of aerogel as an effective pair-breaking agent are now well documented.\cite{halperin}
For the past few years, attention has been shifted to understand phenomena related to an energy scale smaller than the condensation energy. For example, the relative stability among possible superfluid phases, specifically the transition between two superfluid phases observed in this system, the $A$-like and the $B$-like phases, has been investigated. In the absence of a magnetic field, the supercooled $A$-like phase appears at all pressures studied, even below the bulk polycritical point (PCP),\cite{gervais,nazaretski,choithesis} while only a very narrow region where the two phases coexist was identified on warming.\cite{vicente} In the presence of low magnetic fields, the $B$-like to $A$-like transition was observed, on warming, to follow a quadratic field dependence,\cite{brussard,gervais,baumgardner} which is reminiscent of the bulk $A-B$ transition, $1-T_{AB}/T_{c}=g(\beta)(B/B_{c})^{2}$, where $T_{AB}$ and $T_{c}$ are the $A-B$ transition and the superfluid transition temperatures, respectively, and $g(\beta)$ is a strong-coupling parameter that is a function of $\beta$ parameters of the Ginzburg-Landau free energy (see Sec. III). However, the systematic field and pressure dependence study by Gervais {\it et al.\,}\cite{gervais} found a monotonic increase in $g(\beta)$ with pressure without showing any anomalies. This observation raised a question on the position or the existence of the PCP in aerogel. It is important to emphasize that the $A$ and the $B$ phases of bulk $^3$He are highly competing phases separated by first-order transition with a minute-free-energy difference and have identical intrinsic superfluid transition temperatures. These properties are at the heart of many intriguing phenomena showing subtle modifications of the $A-B$ transition in the presence of weak external perturbations such as a magnetic field. Therefore, it is reasonable to expect that the presence of impurities or disorder will have a similar influence on the $A-B$ transition.
In 1996, Volovik\cite{volovik96} discussed the significance of the quenched random anisotropic disorder presented by the strand-like aerogel structure and its interaction with the anisotropic order parameter. This coupling is thought to be particularly important in the $A$-phase, where the order parameter is doubly anisotropic in the sense that the rotational symmetries in spin and orbital space are broken separately. Vicente {\it et al.\,}\cite{vicente} argued that the aerogel strands generated orbital fields emulating the role of a magnetic field, thereby giving rise to similar profound effects on the $A$-like to $B$-like transition. They further suggested the use of uniaxially deformed aerogel to amplify and to systematically investigate the effect of the anisotropic disorder.\cite{vicente} A series of calculations by Aoyama and Ikeda\cite{aoyama,aoyama3} are consonant with these ideas and predict a widened $A$-like phase region in a uniaxially deformed aerogel, the appearance of a novel superfluid phase in uniaxially stretched aerogel, and a change in the PCP location in the phase diagram.
Unlike the $B$-like phase, the clear identification of the $A$-like phase in aerogel has not been made. However, some of the recent NMR measurements using uniaxially deformed aerogels\cite{kunimatsu,elbs} provide compelling evidence that the $A$-like phase possesses the $ABM$ pairing symmetry, albeit with unusual textural configurations. The free-energy calculation by Ikeda and Aoyama\cite{ikeda} also found the disordered $ABM$ phase as the most stable among the various plausible pairing states, such as the Imry-Ma,\cite{volovik06} the planar, and the robust\cite{fomin} phases. Furthermore, the third superfluid phase observed in 98\% aerogel in the presence of high magnetic fields\cite{choia1} fortifies this identification. Therefore, we will continue our discussion with the assumption that the $A$-like phase observed at least in 98\% aerogel has the same pairing symmetry as the bulk $A$-phase.
With this notion, we conducted longitudinal ultrasound attenuation measurements in the superfluid phases of $^3$He in 98\% porosity silica aerogel. Our measurements were performed in the presence of magnetic fields, 0 to 0.444~T, and at various sample pressures ranging from 14 to 33~bar. At the highest field, the existence of the meta-stable $A$-like phase persisted to the lowest temperatures, thereby allowing the sound attenuation in the $A$-like phase to be measured over the entire range of the temperatures studied. In lower magnetic fields, we were able to identify the transitions between the two phases on cooling and warming, and herein, a $P$-$B$-$T$ phase diagram of this system is presented.
\section{Experiment}
The presence of the compliant aerogel complicates the sound propagation because the sound modes of the liquid $^3$He and the aerogel matrix are effectively coupled.\cite{mckenna} As a result, two longitudinal sound modes emerge in this composite medium: one with the speed of sound close to, but slightly lower than, that of the liquid (fast mode) and the other with a significantly lower speed of sound (slow mode).\cite{golov} We measured the longitudinal fast sound attenuation in superfluid $^3$He in 98\% aerogel at frequencies between 3.69 and 11.3~MHz. The employment of the multiple frequency excitations turned out to be extremely valuable in this work for the reason described later in this paper.
Two best-matched LiNbO$_3$ transducers (9.6~mm diameter) with fundamental resonances of 1.1~MHz were selected from six transducers tested using a broadband spectrum analyzer and were used as a transmitter and a receiver. The transducers were supported by a MACOR spacer forming a 3.02~mm size acoustic cavity. Aerogel with 98\% porosity was grown in and around this cavity to ensure optimal acoustic coupling between the aerogel and the transducers. The aerogel grown outside of the cavity was carefully removed, and copper wires were attached to the outer surfaces (electrodes) of the transducers using silver epoxy. In order to reduce the ringing of the transducers, a thin layer of silver epoxy was applied to the electrode. A small piece of a cigarette paper with numerous needle holes was placed between each transducer and the cell wall to interrupt back reflections from the wall through the bulk liquid. The sample cell housing the cavity was placed inside a homemade superconducting solenoid magnet located in the inner vacuum space. The magnet was thermally anchored to the mixing chamber. We chose the magnetic field, $\vec{B}$, to be perpendicular to the sound wave vector $\vec{q}$, $\vec{B}\bot\vec{q}$, expecting ${\vec{l}}\parallel\vec{q}$ in the $A$-like phase, where $\vec{l}$ indicates the orbital angular momentum of the Cooper pair. The top part of the sample cell forms a diaphragm so the pressure of the cell can be measured capacitively. The variation in the cell pressure during the measurement was around $\pm0.1$~bar. A schematic of the experimental geometry can be found elsewhere.\cite{moon}
A commercial spectrometer, LIBRA/NMRKIT II (Tecmag Inc., Houston, TX) was used to transmit 3 $\mu$s pulses and also to detect the transmitted signals. Each measurement was obtained by averaging eight transmitter signals produced in a phase alternating pulse sequence. The level of excitation used in this experiment was set in the range where neither self-heating nor nonlinearity was observed. In one temperature sweep, the measurements at four pre-determined frequencies were performed in a cyclic manner. The temperature was monitored by a melting curve thermometer for $T\geq1$~mK and a Pt-NMR thermometer for $T\leq1$~mK.
In spite of our effort to spoil the quality factor of the transducers, sustained ringings were observed and we were unable to resolve echoes following the initial received signal. Consequently, by integrating a portion of the received signal, only the relative attenuation could be determined. The region of integration was carefully chosen not to include any echoes. Our method produced consistent relative attenuation for various choices of the integration range within the safe window described above. The relative attenuation in reference to the value at the aerogel superfluid transition temperature ($T_{ca}$) was determined by
\begin{equation}
\Delta\alpha=\alpha(T)-\alpha(T_{ca})=-\frac{1}{d}\ln\frac{A(T)}{A(T_{ca})},
\end{equation}
where $d$ is the sound path length and $A(T)$ is the integrated area of the transmitter signal at temperature $T$.
\begin{figure}
\includegraphics[width=3.7in]{figure1.EPS}
\caption{\label{Fig.1.} (Color online) Temperature dependences of relative longitudinal sound attenuations using a 6.22~MHz excitation at 33 bar in the presence of various magnetic fields. All the data were taken on warming after cooling through the $A$-like to $B$-like transition except for $B =$ 0.444~T, where no supercooled transition was observed. The arrows point the positions where the $B$-like to $A$-like phase transitions occur. Inset: expanded view of zero-field attenuation near the superfluid transition indicated by the vertical line.}
\end{figure}
\begin{figure}
\includegraphics[width=3.7in]{figure2.EPS}
\caption{\label{Fig.2.} (Color online) Temperature dependences of relative longitudinal sound attenuations using a 6.22~MHz excitation at 25~bar in the presence of various magnetic fields. See the caption of Fig.\,1 for additional details. Inset: expanded view of zero-field attenuation near the superfluid transition indicated by the vertical line.}
\end{figure}
\section{Results and discussion}
\subsection{Longitudinal sound attenuation and the $A-B$ transition in aerogel}
Figures 1 and 2 show the relative ultrasound attenuations obtained at 33 and 25~bar in the presence of magnetic fields ranging from 0~T to 0.444~T, respectively. All the data shown were taken on warming after cooling though the supercooled $A$-like to $B$-like transition at a fixed external magnetic field, except for $B=$~0.444~T, where no supercooled transition was observed down to $\approx$~200~$\mu$K. Therefore, the warming trace at the highest field should be in the $A$-like phase for the entire temperature range, probably in the meta-stable $A$-like phase in the low temperature region. The superfluid transition is marked by a slight decrease in attenuation around 2.1~mK for 33~bar (Fig.\,1) and 1.9~mK for 25~bar (Fig.\,2). The zero field attenuation, which essentially represents the $B$-like phase attenuation except for a very narrow region ($\approx$~100~$\mu$K) right below $T_{ca}$, can be directly compared with the absolute attenuation measurements by Choi {\it et al.\,}\cite{choi} performed under almost identical experimental conditions. The features observed in the current experiment, namely, the broad shoulder structure appearing in the range $1.0 < T < 1.5$~mK and the absence of attenuation peak associated with the pair-breaking and the order-parameter collective modes, are consistent with those reported earlier\cite{choi} and also with the calculations by a Hiroshima group.\cite{higashitani}
\begin{figure}
\includegraphics[width=3.7in]{figure3.EPS}
\caption{\label{Fig.3.} (Color online) The $A-B$ transition features in sound attenuation at 33 bar. The red or upper (black or lower) trace represents the attenuation in the $A$-like ($B$-like) phase. The top (bottom) panels show the traces taken using 6.22~MHz (8.73~MHz) excitations. The switching behavior between the two phases is clearly demonstrated for each field as marked by an arrow.}
\end{figure}
\begin{figure}
\includegraphics[width=3.7in]{figure4.EPS}
\caption{\label{Fig.4.} (Color online) The $A-B$ transition features in sound attenuation at 25 bar. See the caption of Fig.\,3 for additional details. }
\end{figure}
\begin{figure}
\includegraphics[width=3.7in]{figure5.EPS}
\caption{\label{Fig.5.} (Color online) Temperature dependence of attenuation at 33~bar using 6.22~MHz excitation at two different warming rates. The attenuation in the $B$-like ($B =$~0) and the $A$-like ($B=$~0.444~T) phases are already shown in Fig.\,1. For $B =~$0.111~T, the attenuation was measured with two warming rates of 1.4~$\mu$K (inverted triangles) and 1.7~$\mu$K (triangles). Inset: magnified view of the region of the $A-B$ transition in aerogel.}
\end{figure}
Establishing the attenuation in the $A$-like ($B=$~0.444~T) and the $B$-like ($B=$~0) phases for the entire temperature range in the superfluid, one can envision a transition between the two phases at any intermediate field where a switching from one trace to another occurs. It is expected that the attenuation in the $A$-like phase is higher than in the $B$-like phase under the assumption that it is the $ABM$ state, since the sound presumably propagates along the node direction in our experimental configuration. However, unlike in the bulk, the difference in attenuation between the $A$-like and the $B$-like phases is much smaller and subtle because of the absence of the order-parameter collective modes, which are the fingerprints of specific pairing symmetry, and the presence of the impurity states residing in the gap. One can see the subtle difference in the attenuation between two phases in Figs.\,1 and 2. At all temperatures, the attenuation in the $A$-like phase is slightly larger than in the $B$-like phase, while the largest difference is observed in the zero temperature limit. For this reason, the acoustic signature of the $A-B$ transition in aerogel is not as clear as in the bulk. Despite this small difference in attenuation, the $B$-like to $A$-like transition features are noticeable in most of the cases (indicated by the arrows in Figs.\,1 and 2). However, in the temperature region where two phases show almost identical attenuation, as in $0.7<T<1.0$~mK or very close to $T_{ca}$, the transition feature is rather vague. When this situation arose, the transition temperature $T_{ABa}$ was determined from the attenuation measurements conducted at other frequencies. The magnified views of the $A-B$ transition features are shown in Figs.\,3 and 4 for 33 bar and 25 bar, respectively. For each field, the switching behavior between the $A$-like (red, upper trace) and the $B$-like (black, lower trace) phases is unmistakably demonstrated in these figures. While the transitions at $B=0.333$ and 0.385~T for 33 bar at 6.22~MHz (Fig.\,1) are not clear, the transitions at 8.73~MHz are much more evident in Fig.\,3. This phenomenon is due to the non-trivial frequency dependencies of the attenuation observed in aerogel. The details of this subject are beyond the scope of this paper and will be reported in a separate publication.
\begin{figure}
\includegraphics[width=3.8in]{figure6.EPS}
\caption{\label{Fig.6.}(Color online) Magnetic field dependence of the width of the $A$-like phase, $\Delta T=T_{ca}-T_{ABa}$. For comparison, our results are plotted along with those from Gervais {\it et al.\,} (solid circles)\cite{gervais}. The data points from Gervais {\it et al.\,} were taken at the slightly different pressures of 33.4, 28, 25, and 20~bar, respectively.}
\end{figure}
The lowest finite magnetic field used in this experiment was 0.111~T, and two attenuation measurements performed in this field at 33~bar are shown in Fig.\,5. These data were collected with two different warming rates of 1.4~$\mu$K/min (inverted triangles) and 1.7~$\mu$K/min (regular triangles). Both measurements produced the same transition temperature despite the difference in the warming rate by about 20\%. The small differences between the two traces arise from the background drift associated with the $^4$He bath level and room-temperature variation.
In Fig.\,6, the widths of the $A$-like phase, $\Delta T=T_{ca}-T_{ABa}$, as a function of $B^2$, along with the results obtained in the low-field region by Gervais {\it et al.\,}, are plotted. Within the Ginzburg-Landau (G-L) limit, we can perform analysis that is similar to work used to describe the bulk liquid.\cite{tang} Specifically, the suppression of the $B$-like phase in finite magnetic fields can be written as
\begin{equation}
1-T_{ABa}(T)/T_{ca}=g(\beta)(B/B_{c})^{2}+\mathcal{O}(B/B_{c})^{4}.
\end{equation}
Here, $B_{c}$ represents a characteristic field scale directly related to the transition temperature, namely,
\begin{equation}
B_{c}=\sqrt{\frac{8 \pi^2}{7\zeta(3)}}\frac{k_{B}T_{ca}}{\gamma\hbar}(1+F^{a}_{0}),
\end{equation}
where $k_B$, $\gamma, \zeta(x)$, and $F^{a}_{0}$ are the Boltzmann constant, the gyromagnetic ratio for a $^3$He nuclei, the Riemann zeta function, and a Fermi-liquid parameter, respectively. In addition, the strong-coupling parameter $g(\beta)$ is a function of the pressure-dependent $\beta$-parameters, the coefficients of the quartic terms in the G-L free-energy expansion,\cite{fetter} and can be written as
\begin{multline}
g(\beta)= \frac{\beta_{245}}{2(2\beta_{345}-3\beta_{13})} \times \\
\left(1+\sqrt{\frac{(3\beta_{13}+\beta_{345})(2\beta_{13}-\beta_{345})}{\beta_{245}\beta_{345}}}\right),
\end{multline}
where $\beta_{ijk}=\beta_{i}+\beta_{j}+\beta_{k}$. In the weak coupling limit, $g(\beta)\rightarrow$~1, and the strong-coupling effects cause it to increase.
\begin{figure}
\includegraphics[width=3.7in]{figure7.EPS}
\caption{\label{Fig.7.} (Color online) Magnetic field dependence of the width of the $A$-like phase scaled by $B^2$. The quadratic coefficient, $\textit{g}(\beta)$ is determined by the intersection of the each curve with the $B=$~0 axis, Eq.\,(2).}
\end{figure}
In order to illuminate the overall field dependence, the data presented in Fig.\,6 are recasted as $\Delta T/B^{2}$ in Fig.\,7. As noted by Tang {\it et al.\,},\cite{tang} one of the advantages of this plot is that the intersection of the curve with the $B=0$ axis gives the strong-coupling parameter, $g(\beta)$, and the slope of the curve is related to the coefficient of the higher-order correction, as can be seen in Eq.\,(2). Our $g(\beta)$ values extracted by extrapolating to zero field are shown in Fig.\,8. In the same figure, $g(\beta)$ of the bulk by Tang {\it et al.\,} (open circles) and of 98\% aerogel by Gervais {\it et al.\,} (solid cricles) are included for comparison. Additionally, we reproduced the theoretical calculation\cite{gervais} based on the homogeneous scattering model (HSM)\cite{thunne} with the rescaled strong-coupling corrections by the factor of $T_{ca}/T_c$ for two different mean-free path values of $\ell =$~150 (dotted-dashed line) and 200~nm (dashed line). Although our $g(\beta)$ value at 19.5~bar is in good agreement with that of Gervais {\it et al.\,}, the discrepancy between the two sets of data becomes larger at higher pressures. However, $g(\beta)$ in aerogel from both measurements is substantially smaller than that of the bulk value at the corresponding pressure. For the bulk, $g(\beta)$ grows quickly and approaches the PCP as predicted by the G-L theory. However, no such behavior is seen in aerogel. Although the error bars in our data are rather large, our results lie between the two theoretical curves.
It is also interesting to observe that the sign of the quartic correction is negative at higher pressures and seems to change its sign at $P\approx 19.5$~bar (see Fig.\,7), which needs to be compared with the bulk case where the sign crossover occurs at $P\approx 6.7$~bar.\cite{tang} Based on these observations, one could argue that the presence of aerogel reduces the strong coupling effects and, in effect, the phase diagram of this system is shifted up in pressure.
\begin{figure}
\includegraphics[width=3.7in]{figure8.EPS}
\caption{\label{Fig.8.} (Color online) Pressure dependence of $g(\beta)$. The present data (solid squares) are shown with the data by Gervais {\it et al.\,}(solid circles)\cite{gervais} for aerogel and by Tang {\it et al.\,}(open circles)\cite{tang} for the bulk liquid. The dashed and doted-dashed lines are from HSM with the transport mean-free path, $\ell=$~200 and 150~nm, respectively (see Ref.\,12 for details).}
\end{figure}
\subsection{$A-B$ transition in aerogel by isothermal field sweeps}
The $A-B$ transition can be induced through an isothermal field sweep (IFS). Although it is a time-consuming process, an IFS offers an independent way of determining this phase transition and is especially valuable in the region where the slope of the transition curve in the $T-P$ phase diagram becomes small. During an IFS in either the up or down direction, heating was observed due to the eddy currents in the silver cell body. To alleviate this problem, we slowly demagnetized the main magnet of the nuclear demagnetization stage during a field sweep (typically $\approx 14~\mu$ T/min). This passive procedure limited the temperature variation during an IFS to $\approx$~50~$\mu$K.
In Fig.\,9, the magnitudes of the integrated acoustic signals taken at four different frequencies during an isothermal field sweep at 25~bar and 0.3~mK are displayed. The temperature variation during this process is also shown in the same figure. The sample was cooled from the normal fluid in the presence of a magnetic field of 0.444~T to $\approx$~0.3~mK. After establishing equilibrium, the magnetic field was slowly reduced at the rate of 0.4~mT/min\cite{note1} to go through the $A$-like to $B$-like transition. Therefore, the $B$-like phase was supposed to be induced through a primary nucleation, and this case is the only instance of a primary nucleation transition observed by IFS in our work. For the entire sweep process, the temperature remained within $\approx$~30~$\mu$K around 0.27~mK. The smooth change in magnitudes at all frequencies can be observed from $\approx$~0.43 to 0.4~T, indicating the transition from the $A$-like to $B$-like phase. The difference in the magnitude of the acoustic signal between two phases matches well with the attenuation difference determined from the temperature sweep measurements shown in Fig.\,2.
For $B\lesssim$~0.4~T (in the $B$-like phase), the attenuation exhibits a weak-field dependence, most notably at 11.3~MHz. This behavior can not be simply attributed to the temperature variations during the field sweep because the attenuation shows a very weak temperature dependence around 0.3~mK (see Figs.\,1 and 2). One can speculate that this variation in attenuation might be related to the progressive distortion of the gap induced by magnetic field, as the isotropic $BW$ state evolves through the distorted $BW$ state to the planar state and eventually to the $ABM$ phase with the node along the sound propagation direction.\cite{Teowrdt} The increase (decrease) in the magnitude (attenuation) in the low-field region could be due to the enhancement (reduction) in the component of the gap perpendicular (parallel) to the magnetic field. In the $A$-like phase at the highest field, the sound propagates in the node direction, resulting in a higher attenuation.
\begin{figure}
\includegraphics[width=3.7in, height=3in]{figure9.EPS}
\caption{\label{Fig.9.} (Color online) Results of the IFS at 0.3~mK and $P=$~25~bar. The magnitudes of the integrated acoustic signals, $A(T)$, measured using four different excitation frequencies are displayed as a function of magnetic field. The temperature variation during the IFS is also shown in the bottom panel.}
\end{figure}
Several additional IFS studies were conducted at various combinations of pressure and temperature, where the sample was cooled from the normal state at a fixed field to a temperature in the $B$-like phase via the superfluid and the supercooled $A$-like to $B$-like transitions. Then, the magnetic field was ramped up through the $B$-like to $A$-like transition and decreased again back through the transition, if necessary. Figure 10 shows the IFS results at 14 bar and $T\approx 0.27$~mK. The phase transition occurs over a rather broad range of field ($\Delta B\approx 50$~mT), but no appreciable hysteresis was observed. The results of two other IFS studies at 29~bar ($T\approx$~0.86 and 1.38~mK) are shown in Figs.\,11 and 12. For $T\approx 0.86$~mK (Fig.\,11), the transition can only be identified in the 3.69~MHz measurements ($\Delta B\approx$~20~mT).
\begin{figure}
\includegraphics[width=3.7in]{figure10.EPS}
\caption{\label{Fig.10.} (Color online) Results of the isothermal field sweep at 14 bar.}
\end{figure}
Brussaard {\it et al.\,}\cite{brussard} observed hysteretic behavior in the field driven $A-B$ transition in their measurements at $T \approx$~0.335~mK and $P=$~7.4~bar using an oscillating aerogel sample attached to a vibrating wire. The magnetic field sweep was performed in the presence of a field gradient in which a single A-B phase boundary was moving through the sample during the process. They proposed the pinning of the $A-B$ phase boundary by the aerogel strands as a mechanism for the observed hysteresis. Furthermore, based on this scenario, they made an argument that the $A-B$ transitions determined by a conventional temperature sweep method, specifically those by Gervais {\it et al.\,}, might not provide reliable thermodynamic transition points due to supecooling and superwarming caused by the pinning, suggesting the finite width of the transition is an evidence of the existence of a range of pinning potential strengths.\cite{fishercomment} We would like to point out that the experiments by Gervais {\it et al.\,} and by us were performed without a designed field gradient. In this case, it is also plausible that the random disorder presented by aerogel, more specifically anisotropic disorder, could cause the broadening of the transition.\cite{imry,vicente} The effect of rounding by disorder is also apparent in the superfluid transition, which is a second-order transition and does not involve an interfacial boundary. Imry and Wortis\cite{imry} have made a heuristic argument about the influence of random impurities on a first-order transition. They predicted various degrees of rounding in the transition due to fluctuations (inhomogeneities) of the random microscopic impurities through the simple generalization of the Harris criterion\cite{harris} valid for second-order transition. It is worth noting that the Lancaster group also reported a similar degree of hysteresis in field (approximately millitesla) in the bulk $A-B$ transition induced by a similar method.\cite{bartkowiak} The field sweep performed at 29 bar around 0.86~mK in Fig.\,11 seems to show a glimpse of hysteresis in the 3.69~MHz data. However, we acknowledge that hysteresis at the level of millitesla can not be resolved from our measurements, and the width of the transition is certainly larger than any hysteresis that might exist.
\begin{figure}
\includegraphics[width=3.7in]{figure11.EPS}
\caption{\label{Fig.11.} (Color online) Results of the isothermal field sweep at 29~bar and $T \approx 0.86$~mK.}
\end{figure}
\subsection{Phase diagram}
The $A-B$ transitions in aerogel identified by the temperature sweep at constant field (TSCF) and the IFS are plotted in the $P$-$T$ phase diagram in Fig.\,13. For both methods, the mid-point of the transition in $T$ or $B$ was chosen as the transition point and the actual width of the transition is represented by the error bar. The width in $B$ is translated into the temperature width using the measured field dependence of the $A-B$ transition in aerogel (see Figs.\,1 and 2). The transition points determined by the two different methods exhibit self-consistency within the resolution of our measurements. For example, the IFS transition point at 14~bar was observed at 0.33~T and lies on the extension of the TSCF measurements at 0.333~T, and the 0.37~T IFS point at 29 bar is right on the line for 0.385~T from the TSCF. We could not have obtained the IFS point at 0.421~T and 25~bar by the conventional TSCF at this field.
\begin{figure}
\includegraphics[width=3.7in]{figure12.EPS}
\caption{\label{Fig.12.} (Color online) Results of the isothermal field sweep (ramp up only) at 29~bar and $T \approx 1.38$~mK.}
\end{figure}
The emerging phase diagram, Fig.\,13, from our measurements unambiguously reveals that the $A-B$ phase boundary in 98\% aerogel recedes toward the melting pressure and zero-temperature corner in response to the increasing field. This tendency is robust even when allowing for the possibility of superwarming, which might shift the transition temperature down. This phase diagram is in drastic contrast to that of the bulk.\cite{hahn} First, the slope of the constant-field phase boundary is positive in aerogel but negative in bulk for most of the corresponding pressure range. Second, the phase boundary in the bulk recedes toward $P\approx$~19~bar, which is in close proximity to the bulk PCP, rather than toward the meting pressure. It is noteworthy that the slope of the bulk $A-B$ phase transition line actually changes its sign around the PCP, with a positive slope for $P<P_c$. The observed behavior of the strong-coupling parameter, $g(\beta)$, and these differences can be accounted for qualitatively and naturally by recognizing the reduction in strong-coupling effects due to impurity-scattering.\cite{thunne,schoi,aoyama,aoyama07} Briefly and simply stated, these effects combine to effectively shift the phases and features of the bulk phase diagram up in the pressure to yield the phase diagram for $^3$He in 98\% aerogel.
\begin{figure}
\includegraphics[width=3.7in]{figure13.EPS}
\caption{\label{Fig.13.} (Color online) Phase diagram of superfluid $^3$He in 98\% aerogel. The solid triangles represent the aerogel superfluid transition. The $A-B$ transitions in aerogel obtained by the TSCF are in solid circles and by the IFS in solid stars. The solid lines going through the data points are guide for eyes but conforms to the constant field phase boundaries for 0.111, 0.222, 0.275, 0.333, and 0.385~T, respectively from right to left. For comparison, the constant field $A-B$ phase boundaries for the bulk liquid are shown by the dotted lines\cite{hahn} for 0.1, 0.3, 0.5, 0.55, and 0.58~T, respectively. The numbers right next to the star symbols indicate the mid-field strength of the transition.}
\end{figure}
In G-L theory, the free energy (relative to the normal state) of the $A(B)$ phase is $f_{A(B)} = -\alpha^{2}/{2\beta_{A(B)}}$, where $\alpha = N(0)(T/T_{c}-1)$ is the coefficient of the quadratic term in the G-L free-energy expansion, $N(0)$ is the density of states at the Fermi surface, and $\beta_{A}=\beta_{245}$, $\beta_{B}=\beta_{12}+\beta_{345}/3$. In zero field, the two phases share the same superfluid transition temperature and the PCP is determined by the condition $\beta_{A}(P_c)=\beta_{B}(P_c)$. The presence of a magnetic field introduces an additional term in the G-L expansion given by
\begin{equation}
f_z = g_{z}B_{\mu}A_{\mu i}A_{\nu i}^*B_{\nu} \label{1}.
\end{equation}
Here, $A_{\mu i}$ represents the order parameter of a superfluid state with spin ($\mu$) and orbital ($i$) indices.\cite{vollhardt} The magnetic field couples through the spin channel of the order parameter. With two distinct symmetries in the $A$ and $B$ phase order parameters, this quadratic contribution lifts the degeneracy in the superfluid transition temperature, thereby pushing the $A$-phase $T_{c}$ slightly above that of the $B$ phase. As a result, a narrow region of the $A$-phase must be wedged between the normal and the $B$ phase for $P<P_c$, even for an infinitesimally small magnetic field. The degree of this effect is inversely related to the free-energy difference between two phases, $g(\beta)\propto (\beta_{A}-\beta_{B})^{-1}$, giving rise to the diverging behavior in $g(\beta)$ as $P\rightarrow P_c$.
In the presence of aerogel, the impurity-scattering warrants various corrections to both $\alpha$ and $\beta$ parameters. The first-order corrections obviously come from the suppression of $T_c$ by pair-breaking and incur the reduction in the strong-coupling effects in the $\beta$-parameters simply scaled by $T_{ca}/T_c$. The most extensive calculation of the $\beta$-parameters including various vertex corrections was done by Aoyama and Ikeda.\cite{aoyama07} Their theoretical phase diagram based on those corrections indeed resembles the bulk phase diagram that is, in effect, shifted to lower temperature and, simultaneously, to higher pressure, resulting in the relocation of the PCP to a higher pressure.
Aoyama and Ikeda have also incorporated the anisotropic nature of the aerogel through the angular dependence of the scattering amplitude.\cite{aoyama} In a uniaxially deformed aerogel, the calculation shows the unambiguous effect of global anisotropy as uniform orbital field, represented by an additional quadratic free-energy term,\cite{thunne}
\begin{equation}
f_a = g_{a}a_{i}A_{\mu i}A_{\mu j}^{*}a_{j} \label{2},
\end{equation}
where $\hat{a}$ is a unit vector pointing in the direction of the aerogel strand. The similarity between Eqs.\,(5) and (6) is apparent. The effect of the orbital field produced by the aerogel strands was estimated to be comparable to the effect produced by a magnetic field $\sim$~0.1~T in the case of complete alignment.\cite{vicente} It has been experimentally demonstrated that uniaxial compression indeed induces optical birefringence proportional to the strain and, consequently, global anisotropy into the system.\cite{pollanen,bhupathi}
In a globally isotropic aerogel, however, the local anisotropy comes into play only when $\xi\lesssim\xi_{a}$, where $\xi_{a}$ represents the correlation length of the aerogel and $\xi$ is the pair correlation length.\cite{vicente} In the other limit, the local anisotropy is simply averaged out to produce no effect. As discussed by Vicente {\it et al.\,}, this net local anisotropy should emulate the effect of magnetic field even in the absence of magnetic field in a globally isotropic aerogel. Furthermore, an inhomogeneity in the local anisotropy would cause a broadening of the $A-B$ transition in aerogel in which the mixture of the $A$ and $B$ phases coexists.\cite{imry} Considering $\xi_{a}\approx$~40~-~50~nm in 98\% aerogel, this local anisotropy effect in a globally isotropic aerogel should be more pronounced at higher pressures but is expected to tail off as the pressure decreases to the point where $\xi \sim \xi_{a}$, which occurs around 10~bar. The impressive agreement in $T_{ca}$ between the experiments and the theory of Sauls and Sharma\cite{sauls03} was achieved by incorporating the aerogel correlation length into the depairing parameter of the homogeneous isotropic scattering model.\cite{thunne}
Although the aerogel sample used in this work is supposed to be isotropic, we cannot rule out the possibility of having a weak global anisotropy built into this sample from the sample preparation or the shrinkage occurring during condensation of $^3$He. In either case, the observed behavior in this work as well as others can be explained coherently.\cite{herman,pollanen}
\section{Conclusion}
Longitudinal ultrasound attenuation measurements were conducted in a 98\% uncompressed aerogel in the presence of magnetic fields. Utilizing the metastable $A$-like phase that extended down to the lowest temperature in 0.444~T, we were able to establish the temperature dependence of the attenuation in the $A$-like phase over the entire superfluid region. This arrangement allowed us to determine the $A-B$ transitions in aerogel in various magnetic fields. Based on the transition points on warming, a $P$-$T$-$B$ phase diagram of this system is constructed. The key features of the phase diagram can be understood on the basis of two fundamental points: first, the strong-coupling effect is significantly reduced in this system by impurity-scattering, and second, the anisotropic disorder presented in the form of aerogel strands plays an important role that emulates the effect of a magnetic field.
\begin{acknowledgments}
We wish to acknowledge the support of the NSF under Grants No. DMR-0803516 (Y.L.), No. DMR-0701400 (M.W.M.), No. DMR-0654118 (NHMFL), and the State of Florida.
\end{acknowledgments}
\newpage
|
\section{Introduction}
$n$-point integrals $I_n^R$ in loop momentum space of tensor rank $R$ appear in any realistic evaluation of Feynman diagrams.
There are several ways to calculate them, and one approach expresses them by a small set of scalar integrals.
The first systematic treatment, in a Standard Model calculation, is known as the Passarino-Veltman reduction \cite{Passarino:1978jh} and expresses four-point tensor integrals (and simpler ones) algebraically by scalar one- to four-point functions.
Two of us made use of this scheme in the early days \cite{Fleischer:1980ub,Riemann:1981ga,Mann:1983dv}.
Nowadays tensor reductions became again a topic of research because at LEP2, LHC and ILC the interesting final states typically consist of more than two particles, some of them being massive.
Several dedicated tensor reduction packages have been developed for the calculation of five- and six-point functions.
As open-source packages we like to mention the Fortran packages LoopTools/FF \cite{Hahn:1998yk2,vanOldenborgh:1990yc} (covering $I_n^R$ with $n \leq 5,R \leq 4$) and Golem95 \cite{Binoth:2008uq} (covering $I_n^R$ with $n\leq 6$ and massless propagators) and the Mathematica package hexagon.m \cite{Diakonidis:2008dt,Diakonidis:2008ij} (covering $I_n^R$ with $(n,R) \leq (6,4), (5,3)$).
Here, we describe a recursive implementation for tensor functions $I_n^R$ with $n\leq 6, R\leq n$ for arbitrary internal masses.
We use the Davydychev-Tarasov approach where the tensor integrals are first expressed by scalar integrals with higher dimensions and indices \cite{Davydychev:1991va}.
In a second step, the scalar integrals may be expressed algebraically by scalar one- to four-point functions \cite{Tarasov:1996br} quite similar to the Passarino-Veltman reduction.
In fact, when using the same basis, the approaches are equivalent.
A difference, though, may arise in the algorithmic realization, and as a consequence in the numerical stability and speed of an implementation.
For more comments on the differences of tensor reduction schemes we refer to the literature quoted.
In a recent paper \cite{Diakonidis:2009fx}, we introduced a convenient and easy-to-program version of the reduction a la Davydychev-Tarasov, which allows a recursive determination of a chain of tensors.
Here we will describe that scheme and present some numerical results.
The integrals to be evaluated are:
\begin{eqnarray}\label{definition}
I_n^{\mu_1\cdots\mu_R} &=& ~ C(\varepsilon) ~\int \frac{d^d k}{i\pi^{d/2}}~~\frac{\prod_{r=1}^{R} k^{\mu_r}}{\prod_{j=1}^{n}c_j^{\nu_j}},
\end{eqnarray}
where the denominators $c_j$ have \emph{indices} $\nu_j$ and \emph{chords}
$q_j$:
\begin{eqnarray}\label{propagators}
c_j &=& (k-q_j)^2-m_j^2 +i \epsilon .
\end{eqnarray}
The normalization $C(\varepsilon)$ plays a role for divergent integrals only and is conventional:
\begin{eqnarray}
C(\varepsilon) &=& (\mu)^{2\varepsilon}~\frac{\Gamma(1 - 2\varepsilon)}{\Gamma(1 + \varepsilon) \Gamma^2(1 - \varepsilon)}.
\end{eqnarray}
Here, we use $d=4-2\epsilon$ and $\mu=1$.
For the evaluation of the scalar functions we will rely on either LoopTools or QCDloops/FF \cite{Ellis:2007qk,vanOldenborgh:1990yc}, and the latter one uses also $C(\varepsilon)$ as defined here.
\section{Recursions}
Our recursions begin with six-point functions where the well known formula \cite{Fleischer:1999hq,Binoth:2005ff,Denner:2005nn,Diakonidis:2008ij} may be used:
\begin{eqnarray}\label{tensor6general}
I_6^{\mu_1 \dots \mu_{R-1} \rho} =
- \sum_{s=1}^{6}
I_5^{\mu_1 \dots \mu_{R-1} ,s } \bar{Q}_s^{\rho}.
\end{eqnarray}
The auxiliary vectors $\bar{Q}_s^{\rho}$ read:
\begin{eqnarray}
\bar{Q}_s^{\rho}&=&\sum_{i=1}^{6} q_i^{\rho} \frac{{0s\choose 0i}_6}{{0\choose 0}_6}~~~,~~~ s=1 \dots 6.
\label{Q6}
\end{eqnarray}
The $I_{n-1}^{\{\mu_1,\cdots\},s}$ is obtained from $I_{n}^{\{\mu_1,\cdots\}}$ by shrinking line $s$, and the ${i,j,\cdots \choose k,l,\cdots}_n$ are signed minors of the modified Cayley determinant ${\choose}_n$ \cite{Melrose:1965kb}.
For further details of notations we refer to \cite{Diakonidis:2009fx}.
\begin{figure}[tb]
\begin{center}
\rotatebox{-90}{\includegraphics*[angle=0,height=15.0cm]{the-triangle}}
\caption{\label{fl-triangle}The recursion triangle.}
\end{center}\end{figure}
The further calculational chain may be read off from figure \ref{fl-triangle}.
Its basic idea is to represent an $n$-point tensor $I_n^R$ by an $n$-point tensor of lower rank $I_n^{R-1}$ and by all the $(n-1)$-point tensors of lower rank $I_{n-1}^{R-1}$.
\footnote{Similar recursive realizations of the Passarino-Veltman reduction may be found in \cite{Denner:2005nn}, see there figures~ 2 and~ 3.}
For 5-point functions, we derived in \cite{Diakonidis:2009fx}:
\begin{eqnarray}\label{tensor5general}
I_5^{\mu_1 \dots \mu_{R-1} \mu} &=&I_5^{\mu_1 \dots \mu_{R-1}} Q_0^{\mu} - \sum_{s=1}^{5}
I_4^{\mu_1 \dots \mu_{R-1},s } Q_s^{\mu},
\\
\label{Qs}
Q_s^{\mu}&=&\sum_{i=1}^{n} q_i^{\mu} \frac{{s\choose i}_n}{\left( \right)_n},~~~ s=0, \dots, n.
\end{eqnarray}
The formula is the analogue to (\ref{tensor6general}).
For $n$-point functions with $n<5$, the corresponding representations contain additional terms because the number of independent chords is then less than four so that the chords don't form a complete basis for $d=4$.
There are several modifications to be applied, and we like to reproduce only one example with auxiliary terms:
\begin{eqnarray}
\label{tensor44}
I_4^{\mu \nu \lambda \rho} &=&I_4^{\mu \nu \lambda} Q_0^{\rho} - \sum_{t=1}^{4}
I_3^{\mu \nu \lambda,t } Q_t^{\rho}
-
G^{\mu \rho } T^{\nu \lambda }- G^{\nu \rho } T^{\mu \lambda }-G^{\lambda \rho } T^{\mu \nu },
\end{eqnarray}
with the additional tensor and vector components:
\begin{eqnarray} \label{eq03}
T^{\mu \nu }&=&I_4^{\mu,[d+]} Q_0^{\nu} - \sum_{t=1}^{4} I_3^{\mu,[d+],t} ~ Q_t^{\nu}~-G^{\mu \nu }I_{4}^{[d+]^2} ,
\\
\label{Gml}
G^{\mu \lambda}&=&\frac{1}{2} g^{\mu \lambda}-
\sum_{i,j=1}^{4} q_i^{\mu} q_j^{\lambda} \frac{{i\choose j}_4}{\left( \right)_4},
\\
I_4^{\mu,[d+]}
&=& I_4^{[d+]}
Q_0^{\mu} - \sum_{t=1}^{4} I_3^{[d+],t} Q_t^{\mu},
\label{GV}
\\
I_3^{\mu,[d+],t}
&=&
I_3^{[d+],t} Q_0^{t,\mu} - \sum_{u=1}^{4} I_2^{[d+],tu} Q_u^{t,\mu},
\label{Wt}
\\\label{Qst}
Q_u^{t,\mu}&=&\sum_{i=1}^{4} q_i^{\mu} \frac{{ut\choose it}_4}{ {t \choose t}_4},~~~ u=0, \dots, 4.
\end{eqnarray}
They may be, finally, represented by the scalar integrals in $d$ dimensions $I_{4},I_{3 }^{t},I_{2 }^{tu}, I_{1}^{tuw}$, where the indices $t,u,w$ indicate truncations of corresponding lines:
\begin{eqnarray}
\label{eq04}
I_{4}^{[d+]^2}&=&\left[\frac{{0\choose 0}_4}{\left( \right)_4} I_{4}^{[d+]} -
\sum_{t=1}^{4} \frac{{t\choose 0}_4}{\left( \right)_4} I_{3}^{[d+],t} \right]{\frac{1}{d-1}},
\\
I_{4}^{[d+]}&=& \frac{{0\choose 0}_4}{{\choose }_4} I_{4}
-
\sum_{t=1}^{4} \frac{{t\choose 0}_4}{{\choose }_4} I_{3}^{t},
\\
I_{3 }^{[d+],t}&=& \left[
\frac{{0t\choose 0t}_4}{{t\choose t}_4}I_{3 }^{t}-
\sum_{u=1}^4\frac{ {ut\choose 0t}_4}{{t\choose t}_4} I_{2 }^{tu} \right]{\frac{1}{d-2}} ,
\label{A301}
\\
\label{eq05}
I_{2}^{[d+],tu}&=&\left[\frac{{0tu\choose 0tu}_4}{{tu\choose tu}_4} I_{2}^{tu}-
\sum_{w=1}^{4} \frac{{0tu\choose wtu}_4}{{tu\choose tu}_4} I_{1}^{tuw} \right]{\frac{1}{d-1}}
.
\end{eqnarray}
The representations for the simpler tensors have been given in \cite{Diakonidis:2009fx}.
\section{Numerical results}
\begin{table}[tb]
\centering
\begin{tabular}{|l|r@{.}l|r@{.}l|r@{.}l|r@{.}l|}
\hline
$p_1$ & 0&5 & 0&0 & 0&0 & 0&5 \\
$p_2$ & 0&5 & 0&0 & 0&0 &-- 0&5 \\
$p_3$ & -- 0&19178191&-- 0&12741180 &-- 0&08262477 &-- 0&11713105 \\
$p_4$ & -- 0&33662712 & 0&06648281 & 0&31893785 & 0&08471424 \\
$p_5$ & -- 0&21604814 & 0&20363139 &-- 0&04415762 &-- 0&05710657
\\ \hline
\multicolumn{9}{|c|}{$p_6 =-(p_1+p_2+p_3+p_4+p_5)$}
\\
\hline
\end{tabular}
\caption{\label{kinem-hexagon-massless}Phase space point of massless six-point functions taken from \cite{Binoth:2008uq}. }
\end{table}
An example of a completely massless tensor reduction uses the momenta given in table \ref{kinem-hexagon-massless}.
We combined our tensor reduction with the scalar master integrals from QCDloop \cite{Ellis:2007qk}.
Table \ref{sixpoint-golem} contains sample tensor components of a six-point function with rank $R=5$.
It shows an agreement of eight digits between the results of our package Hexagon.F \cite{diakonidis-hexagonV0.9:2009} and those of Golem95 \cite{Binoth:2008uq} for the constant terms of the tensor components.
\begin{table}[htb]
\begin{center}
\begin{tabular}{|l|r@{.}l r@{.}l | r@{.}l r@{.}l |}
\hline
& \multicolumn{4}{|c|}{Hexagon.F} & \multicolumn{4}{|c|}{Golem95}
\\
\hline
${F^{03121}}$ & 0&158428987E+0,& 0&41670698E--1 & 0&158428981E+0 ,& 0&41670700E--1
\\
\hline
${F^{11020}}$
& -- 0 & 143913860E+1,& -- 0&16464705E+0 & -- 0&143913853E+1 ,& -- 0&16464708E+0\\
\hline
${F^{20200}}$
& 0&242928780E+2,& 0&55504184E+2 & 0&242928776E+2 ,& 0&55504182E+2
\\
\hline
${F^{22130}}$
& 0&225563941E+0,& 0&23192857E+0 & 0&225563949E+0 ,& 0&23192851E+0
\\
\hline
${F^{33333}}$
& 0&244568135E+0,& 0&74014604E+0 & 0&244568138E+0 ,& 0&74014610E+0\\
\hline
\end{tabular}
\end{center}
\caption[]{\label{sixpoint-golem}Real and imaginary parts of selected tensor components of rank $R=5$ massless hexagon integrals;
comparison of the packages Hexagon.F and Golem95.}
\end{table}
\begin{figure}[t]
\begin{center}
\includegraphics[scale=1.1]{6pt}
\caption{\label{fig:6ptfig}
Momenta flow for the massive six-point topology.}
\end{center}
\end{figure}
As a second example, we reproduce components of the massive tensor integral $I_6^{\alpha\beta\gamma\delta\epsilon}$ in table \ref{sixpoint-massive}.
The kinematics is defined by figure \ref{fig:6ptfig} (with $q_0=0$) and table \ref{kinem-hexagon-massive}.
All tensor components are finite.
The numbers could not be checked by another open source program, but we had the opportunity to compare them with an unpublished numerical package \cite{Uwer-privcommun:2009}.
\begin{table}[th]
\centering
\begin{tabular}{|r|r|r@{.}l|r@{.}l|r|}
\hline
$p_1$ & 0.21774554~E+03 & 0&0 & 0&0 & 0.21774554~E+03 \\
$p_2$ & 0.21774554~E+03 & 0&0 & 0&0 & -- 0.21774554~E+03\\
$p_3$ & -- 0.20369415~E+03 & -- 0&47579512~E+02 &0&42126823~E+02 &0.84097181~E+02 \\
$p_4$ & -- 0.20907237~E+03 & 0&55215961~E+02 &-- 0&46692034~E+02& -- 0.90010087~E+02 \\
$p_5$ & -- 0.68463308~E+01& 0&53063195~E+01 & 0&29698267~E+01 & -- 0.31456871~E+01 \\
$p_6$ & -- 0.15878244~E+02 & -- 0&12942769~E+02 & 0&15953850~E+01 & 0.90585932~E+01
\\
\hline
\multicolumn{7}{|c|}{$m_1 = m_2 = m_3 = m_5 = m_6 = 110.0, ~~m_4 = 140.0$}
\\
\hline
\end{tabular}
\caption{\label{kinem-hexagon-massive}Randomly chosen phase space point of six-point functions with massive particles.}
\end{table}
\begin{table}[htb]
\begin{center}
\begin{tabular}{|c|r@{.}l r@{.}l|}
\hline
& \multicolumn{4}{|c|}{Hexagon.F}
\\
\hline
${F^{03121}}$ & 0&29834730E--09, & -- 0&68229122E--10
\\
\hline
${F^{11020}}$ & 0&42830755E--09, & 0&42574811E--09
\\
\hline
${F^{20200}}$ & -- 0&71172947E--08, & 0&10102923E--07
\\
\hline
${F^{22130}}$ & -- 0&29200434E--09, & 0&78553811E--10
\\
\hline
${F^{33333}}$ & 0&17451484E--07, & -- 0&30914316E--07
\\
\hline
\end{tabular}
\end{center}
\caption[]{\label{sixpoint-massive}Selected tensor components of rank $R=5$ massive hexagon integrals;
}
\end{table}
Further numerical results may be found in the transparencies of the talk \cite{Riemann-radcor:2009}.
\section*{Acknowledgments}
Work supported in part by Sonderforschungsbereich/Trans\-re\-gio SFB/TRR 9 of DFG
``Com\-pu\-ter\-ge\-st\"utz\-te Theoretische Teil\-chen\-phy\-sik"
and by the European Community's Marie-Curie Research Trai\-ning Network
MRTN-CT-2006-035505
``HEPTOOLS''.
J.F. likes to thank DESY for kind hospitality.
\input{radcor2009-hep-ph-riemann.bbl}
\end{document}
|
\section{Introduction}
The microscopic mechanism that generates high temperature superconductivity in the cuprates continues to be controversial.
One set of proposals is based on the analogy with heavy fermion metals where a superconducting dome is observed surrounding
the quantum critical point (QCP) that arises as antiferromagnetism is suppressed by an external parameter such as pressure.\cite{mat}
In this case the pairing glue arises from the exchange of the soft longitudinal antiferromagnetic fluctuations
in the vicinity of the QCP. In the cuprates doping plays the role of the external parameter and there are several proposals
for the nature of the QCP that appears near optimal doping involving fluctuations in various order parameters e.g. nematic,\cite{kivel}
d-density wave\cite{chak} and orbital currents\cite{var} in addition to antiferromagnetism.\cite{sach} A second set goes back to
Anderson's very early proposal
that the strong singlet nearest neighbor correlations in the 2-dimensional Heisenberg antiferromagnet generates pairing
when doped holes are introduced. The advocates of this resonant valence bond (RVB) mechanism point to the strong
asymmetry in the cuprate phase diagram between the physical behavior on the under- and overdoped sides of optimal
doping and the QCP. This contrasts strongly with the symmetric dome observed in heavy fermions. Further the highly
anomalous physical properties that characterize the pseudogap phase at underdoping are associated with a short range
spin liquid in the cleanest cuprate materials, e.g. YBa$_2$Cu$_4$O$_8$
and HgBa$_2$CuO$_{4+x}$. Nonetheless strong correlations
and the absence of a broken translational symmetry in the pseudogap phase have proved to be formidable obstacles to
constructing a comprehensive microscopic RVB theory for underdoped cuprates. For more
details see several recent reviews.\cite{pwa,gros,leeP,ogfu}
Several years ago we proposed a 2-dimensional array of weak coupled 2-leg Hubbard ladders as an example of a model
where occurs a truncation of the full Fermi surface to pockets associated with hole or electron doping in a system without
broken symmetry.\cite{konrice} Subsequently this model led to a phenomenological ansatz for the propagator in underdoped
cuprates starting from a renormalized mean field description of an undoped RVB spin liquid insulator.\cite{rice1} This
phenomenological propagator has been recently used successfully to fit a range of experiments covering many anomalous
properties of the pseudogap phase.\cite{rice2,pdj,val,car1,car2,car3,car4}
In this paper we extend our earlier analysis to the case of an array of lightly doped 4-leg Hubbard ladders with an onsite
weak interaction. Our goal is to construct a tractable 2-dimensional model with a partially truncated Fermi surface in
which d-wave pairing arises on the residual Fermi surface through scattering in the Cooper channel.
Earlier numerical renormalization group studies on the 2-dimensional Hubbard model were interpreted as pointing towards a
similar pairing mechanism.\cite{Honerkamp,Laeuchli} A key feature of the present model is the presence of a finite energy
Cooperon resonance in the pseudogap which is generated in association with the partial truncation of the Fermi surface. D-wave
pairing follows on the remnant Fermi surface through the coupling to the Cooperon.
\section{Four Leg Hubbard Ladders}
The properties of a single 4-leg Hubbard ladder with open boundaries have been studied extensively in
both the weak and strong coupling limits. We consider here the former with equal nearest neighbor
hopping $t_{0}$, along the legs and rungs. In this case the 4 bands split into two band pairs.
The inner pair, $A_{1,2}$, are standing waves on the rungs with wavevectors $(2\pi /5,3\pi /5)$.
At half filling the corresponding Fermi wavevectors are $k_{FA_1} = \pm 3\pi /5 $ and $k_{FA_2} = \pm 2\pi /5$
leading to a common Fermi velocity,
$v_{FA} = 2t_{0} \sin(2\pi /5)$. The outer band pair, $B_{1,2}$, have Fermi wavevectors $K_{FB_2}= \pm\pi/5$
and $K_{FB_1} =\pm 4\pi /5$ and a smaller Fermi velocity, $v_{FB} =2t _{0}\sin(\pi /5)$.
We obtain a band structure of four bands with energies
\begin{eqnarray}\label{eIIi}
E_{A_{1,2}}(k) &=& \epsilon_{\parallel}(k) \mp 2t_0\cos(2\pi/5), \cr
E_{B_{1,2}}(k) &=& \epsilon_{\parallel}(k) \mp 2t_0\cos(\pi/5),
\end{eqnarray}
where $\epsilon_{\parallel}(k)$ represents the dispersion along the ladder.
The annihilation (creation) operators of electrons of the outer and inner bands,
denoted as $B_{1,2}, B^\dagger_{1,2}$ and $A_{1,2}, A^\dagger_{1,2}$ respectively, are
\begin{eqnarray}\label{eIIii}
B_1 &=& \sum_{n=1}^4\sin( \pi n/5)c_n;\cr\cr
B_2 &=& \sum_{n=1}^4\sin(4\pi n/5)c_n;\cr\cr
A_1 &=& \sum_{n=1}^4\sin(2\pi n/5)c_n;\cr\cr
A_2 &=& \sum_{n=1}^4\sin(3\pi n/5)c_n,
\end{eqnarray}
where $c_n$ is the corresponding annihilation operator of an electron on the n-th leg
of the ladder.
\begin{figure}
\begin{center}
\begin{tabular}{cc}
\epsfig{file=4leg.eps,width=0.16\linewidth,clip=}&
\hskip.4in\epsfig{file=bandstr.eps,width=0.29\linewidth,clip=}\\
\end{tabular}
\end{center}
\caption{On the l.h.s. of the figure is picture a four leg ladder with equal hopping along
and between the legs of the ladder. On the r.h.s are pictured the corresponding four bands, $A_{1,2}$, $B_{1,2}$ of such a ladder.}
\label{bs}
\end{figure}
Close to half filling the Fermi velocities of the outer band pair labeled by $B$ are smaller than those of the inner bands
labeled by $A$, so that in the presence of interactions the effective dimensionless coupling constants for electrons
in the inner bands are smaller than those for the outer bands.
In the weak coupling limit, i.e. an onsite interaction characterized by $U \ll t$, this Fermi
velocity difference leads to a large difference in the characteristic energy scales and to a decoupling of the RG flows of the two band pairs.
The outer band pair has the larger critical energy scale and flows to strong coupling first as
the energy scale is lowered.\cite{Ledermann,Affleck,LeHur} The inner band pair has a lower critical scale.
Therefore in the first approximation one can treat inner and outer bands of individual 4-leg
ladders as decoupled from each other. Then each band pair will effectively constitute a two-leg ladder.
It is well known that two-leg ladders acquire spectral gaps for quite general interaction patterns.
For the inner bands the smaller dimensionless couplings lead to smaller spectral gaps.
At half filling each band pair is exactly half filled and behaves as a half filled 2-leg Hubbard ladder.
The difference in the energy scales leads to a finite doping range $x < x_c$ where all the doped holes
enter the inner band pair and the outer band pair remains exactly half filled. We note in passing that
similar behavior is found also in the strong coupling limit, $U \gg t$.\cite{EsTs}
Given that a 4-leg ladder can be reduced to two 2-leg ladders, we will now recall some basic facts about 2-leg ladders.
For general interactions they become either Luttinger liquids or dynamically generate spectral gaps.
In the latter case an increased symmetry appears at small energies where a half filled 2-leg ladder can
be well described by the O(8) Gross-Neveu model.\cite{so8}
The Gross-Neveu model is exactly solvable for all
semi-simple symmetry groups and a great deal is known about its thermodynamics and correlation functions.
In the SO(8) case the correlation functions were studied in Refs. (\onlinecite{KonLud,EsKon}).
Since the model itself has Lorentz symmetry, all excitation branches have relativistic dispersion laws:
\begin{equation}\label{eIIiii}
E(p) = \sqrt{(vp)^2 +M^2}.
\end{equation}
The spectrum consists of three octets of particles of mass $\Delta$ and a
multiplet of 29 excitons with mass $\sqrt 3 \Delta$.
Two octets consist of quasi-particles of different chirality transforming according to the
two irreducible spinor representations of SO(8), while the third octet consists of vector particles.
The latter include magnetic excitations as well as
the Cooperon (a particle with charge $\pm 2e$).
The 16 kink fields, carrying charge, spin, orbit, and parity indices, are direct descendants
of the original electron lattice operators on the ladders.
The SO(8) GN model describes several different phases related to one another by particle-hole transformations.
Which phase is realized depends on the bare interaction. In this paper we assume that it is in the so-called
D-Mott phase (in the terminology of Ref. \onlinecite{so8}).
On the two two-leg ladders ($A$ and $B$), the superconducting (SC) order parameters are given by
\begin{eqnarray}\label{eIIiv}
\Delta_{A} &=& A_{1,\uparrow}A_{1,\downarrow} - A_{2,\uparrow}A_{2,\downarrow} ;\cr
\Delta_{B} &=& B_{1,\uparrow}B_{1,\downarrow} - B_{2,\uparrow}B_{2,\downarrow} .\label{OP}
\end{eqnarray}
The distinct feature of the half filled ladder is that this order parameter is purely real and has a Z$_2$ symmetry.
However, the symmetry is restored to U(1) and the phase stiffness becomes non-zero as soon as doping
is introduced. It is an interesting feature of the SO(8) GN model that the only mode which becomes
gapless at finite doping is the Cooperon. Neither magnetic excitations, nor quasi-particles become
gapless.\cite{evans} When the doping increases the SO(8) GN model gradually crosses over to the
SO(6) GN one plus the U(1) Gaussian model. The latter model describes the fluctuations of the
superconducting phase. The effective low energy bosonized Lagrangian density
for the Cooperon field, $\Phi$, is
\begin{equation}\label{eIIv}
{\cal L} = \frac{K}{8\pi}[v^{-1}(\partial_{\tau}\theta)^2 + v(\partial_x\theta)^2], ~~ \Phi = \Delta_0\mbox{e}^{i\frac{\phi}{2}}
\end{equation}
where $\phi$ is the field dual to $\theta$. (Here
-- according to Ref. (\onlinecite{huber}) -- the Luttinger parameter $K$ depends weakly on doping and is always in the range $1 > K > 0.9$.
On the other hand, the phase velocity is strongly doping dependent.)
For values of doping close to the Cooperon band edge ($|\mu - \Delta/2| \gg \Delta$) spectral curvature is important and
the action given in Eqn. \ref{eIIv} is inadequate. A better description of the Cooperon dynamics is given
by the sine-Gordon model
\begin{equation}\label{eIIvi}
{\cal L} = \frac{1}{8\pi}\left[v_F^{-1}(\partial_{\tau}\theta)^2 + v_F(\partial_x\theta-4\mu)^2\right]
- \frac{M}{2}\cos(\theta).
\end{equation}
where $M^2 = \Delta^2-4\mu^2$.
The mass term here can be thought to arise as follows in a mean field way from the SO(8) Gross-Neveu model. The SO(8) Gross-Neveu
model can be written in terms of fundamental fermions (which are non-local with respect to the original
fermions in the problem) with an interaction term of the form
\begin{equation}\label{eIIvii}
H^{SO(8)}_{int} = 2g(\sum_a\psi^\dagger_a\tau^y\psi_a)^2
\end{equation}
Here $a=1,4$ and $\psi_a = (\psi^R_a,\psi^L_a)$ and $\tau^y$ is a Pauli matrix acting in $R-L$ space. The four fundamental
fermions correspond to the different degrees of freedom in SO(8): charge, spin, orbital, parity. The Cooperon (charge)
we take to be given by $\psi_1$. With a finite chemical potential lowering the Cooperon gap, the fluctuations of the Cooperon
will be strongest. Invoking mean field theory, we thus replace $\psi^\dagger_a\tau^y\psi_a$ for $a=2,3,4$ by its expectation value.
The resulting bosonization of the remaining degree of freedom $\psi_a$ results in the sine-Gordon model.
\section{Superconductivity of Arrays of Four-Leg Ladders: Two Scenarios}
Having elucidated the properties of individual 4-leg ladders, we now consider an array of such ladders.
We assume initially that the electron-electron interaction acts only inside individual ladders and is
much smaller than the bandwidth $W \sim 2t_0$. It is also assumed that $W \gg t_{\perp}$ (the inter-ladder tunneling).
We imagine two scenarios. In the first we assume $t_\perp$ is on the same order as $\Delta_A$, the gap on the inner bands
of the four leg ladder, but much smaller than $\Delta_B$, the gap on the outer bands. In this case coupling the ladders together
lead to small Fermi pockets, very much like in Ref.(\onlinecite{konrice}). However in this case the pockets are found
near $\pm\pi/2,\pm\pi/2$. The residual coupling between these Fermi pockets and
the A-cooperons then leads to superconductivity in the A-bands. And because of a proximity effect, the superconductivity
of the A-bands induces superconductivity in the B-bands.
In the second scenario, we assume $\Delta_A \ll t_\perp \ll \Delta_B$. In this case $t_\perp$ wipes out the effects of interactions
on the A-bands. Coupling them together then gives us an anisotropic two dimensional Fermi liquid. But as $t_\perp$ is much
smaller than $\Delta_B$, the Cooperons on the outer bands at zeroth order remain unperturbed. The coupling then between
the anisotropic Fermi liquid and the B-Cooperons induces superconductivity in the system as a whole. This superconductivity
is d-wave in nature.
We now elaborate on these two scenarios.
\subsection{Scenario I}
We treat the interladder hopping through a random phase approximation (RPA) analysis of the interladder hopping.
The form of the hopping is taken to be long range
\begin{equation}\label{eIIIi}
H_{\rm interladder} = -\sum_{n\neq m,a,b} t^{n,m}_{a,b}c^\dagger_{n,a}c_{m,b},
\end{equation}
where $a,b=1,...,4$ run over the legs of an individual ladder and $n$ and $m$ mark the $n$'th and $m$'th ladders.
By particle-hole symmetry the hopping is assumed to have peaks both near $k_\perp = 0$ and $k=G/2$ where $G=(0,\pi/2)$ is the inverse lattice
vector perpendicular to the ladders. In particular the hopping takes the form
\begin{equation}\label{eIIIii}
t^{n,m}_{a,b} = (1-(-1)^{n-m})f_{ab}(m-n)
\end{equation}
where $f_{ab}(m-n)=f_{ba}(n-m)$ and $f_{ab}(0)= 0$ (i.e. no (additional) hopping within a ladder).
\begin{figure}
\centering
\epsfig{file=4legarray.eps,width=0.4\linewidth,clip=}
\caption{An array of four leg ladders. As an example of the hopping assumed in the RPA
analysis (hopping between every second leg), we show how electrons can hop between the fourth
chain of the n-th 4-leg ladder and the chains on the $n+1$-th ladders.}
\end{figure}
By treating $H_{\rm interladder}$ in an RPA approach, we find that the
single particle Green's function takes the form
\begin{eqnarray}\label{eIIIiii}
G^{\rm 2D~RPA}_{\rm ret}(\omega,k_x,k_y)
&=& \Gamma_1(k_y) G^{\rm 2D}_{A_1}(\omega,k_x,k_y) + \Gamma_2(k_y) G^{\rm 2D}_{A_2}(\omega,k_x,k_y);\cr\cr
G^{\rm 2D}_{A_i}(\omega,k_x,k_y)
&=& \frac{G_{A_i}(\omega,k_x)}{1 + G_{A_i}(\omega,k_x)t^{eff}_i(k_y)},
\end{eqnarray}
where
\begin{eqnarray}\label{eIIIiv}
t^{eff}_1(k_y) &=& 2\sum_{n>0}\cos(4k_y) \cr\cr
&& \hskip -.5in \times \bigg(2(s_1^2+s_2^2)t^{n,0}_{1,1}+(2s_1s_2-s_1^2)(t^{n,0}_{1,2}+t^{n,0}_{2,1})-2s_1s_2
(t^{n,0}_{3,1}+t^{n,0}_{1,3}) - s_2^2 (t^{n,0}_{1,4}+t^{n,0}_{4,1})\bigg);\cr\cr
t^{eff}_2(k_y) &=& 2\sum_{n>0}\cos(4k_y) \cr\cr
&& \hskip -.5in \times \bigg(2(s_1^2+s_2^2)t^{n,0}_{1,1}-(2s_1s_2-s_1^2)(t^{n,0}_{1,2}+t^{n,0}_{2,1})-2s_1s_2
(t^{n,0}_{3,1}+t^{n,0}_{1,3}) + s_2^2 (t^{n,0}_{1,4}+t^{n,0}_{4,1})\bigg);\cr\cr
\Gamma_1(k _y) &=& 2(s_1^2+s_2^2)+2(2s_1s_2-s_1^2)\cos(k_y)
-4s_1s_2\cos(2k_y)-2s_2^2\cos(3k_y);\cr\cr
\Gamma_2(k_y) &=& 2(s_1^2+s_2^2)-2(2s_1s_2-s_1^2)\cos(k_y) -4s_1s_2\cos(2k_y)+2s_2^2\cos(3k_y),
\end{eqnarray}
and $s_1 = \sin(\pi/5)$ and $s_2 = \sin(2\pi/5)$.
We have assumed the
hopping is real and that the low energy contribution to $G^{\rm
2D~RPA}$ comes from the $A$-bands as $\Delta_A \ll \Delta_B$.
Thus $G_{A1}(\omega,k_x)/G_{A2}(\omega,k_x)$ are the Green's functions of the $A$-band
electrons on a given 4-leg ladder. As we have discussed in the
previous section $G_{A1}/G_{A2}$ are no more than the
bonding/anti-bonding electron Greens functions for a 2-leg ladder.
The RPA does not mix $G_{A1}$ and $G_{A2}$ as the weights
of the two are found near differing Fermi wavevectors (i.e.
we can take $G_{A1}(k)G_{A2}(k) \sim 0$ safely for all $k$).
The presence of $\Gamma_1(k_y)$ and $\Gamma_2(k_y)$ act
as structure factors which cause the quasiparticle weight at
various $k_y$ to be negligible. While the denominator
of $G^{\rm 2D~RPA}$ has the periodicity of the reduced Brillouin zone
i.e. $k_y$ and $k_y+\pi/2$ are identified)
these structure functions merely have the periodicity of the original zone
i.e. $k_y$ and $k_y+2\pi$ are identified).
The Green's functions for $A_1/A_2$ at zero chemical potential are given by
\begin{equation}\label{eIIIv}
G_{A_i}(\omega,k_x) = Z_i \frac{\omega + E_{A_i}(k_x)}{\omega^2-E^2_{A_i}(k_x)-\Delta_A^2}
\end{equation}
where the $E_{A_i}$ are defined in Eqn. \ref{eIIi}.
At a chemical potential, $\mu$, that does not exceed the gap,
$G_{A_i}$ is given by
$G_{A_i}(\omega,\mu,k) = G_{A_i}(\omega-\mu,0,k)$
\begin{figure}
\centering
\begin{tabular}{cc}
\epsfig{file=fs.eps,width=0.4\linewidth,clip=}&
\epsfig{file=fs1.eps,width=0.4\linewidth,clip=}\\
\end{tabular}
\caption{The electron and hole pockets of an array of weakly coupled
four leg ladders shown in a periodic zone scheme. On the l.h.s.
of the figure are pockets at zero chemical potential. On the r.h.s. of the figure are pictured the
pockets for finite chemical potential such that the interladder hopping satisfies
$2\Delta_A+2\mu> |t_0| > 2\Delta_A-2\mu$.}
\end{figure}
The excitations are then given by the locations of the poles in
$G^{\rm 2D~RPA}$. These poles then imply that the excitations have the
dispersion
relation
\begin{equation}\label{eIIIvi}
E_i(k_x,k_y) = \mu - \frac{t^{eff}_i(k_y)}{2} \pm \sqrt{(E_{A_i}(k_x)-t^{eff}_i(k_y)/2)^2+\Delta_A^2}.
\end{equation}
For sufficiently large $t^{eff}_i$ a Fermi surface forms (found by
solving $E_i = 0$) consisting of electron and hole pockets. The type
of
pocket is determined by the sign of the effective hopping
\begin{eqnarray}\label{eIIIvii}
t^{eff}_i(k_y) > 2\Delta_A + 2\mu &\rightarrow& {\rm electron~pocket};\cr
t^{eff}_i(k_y) < -2\Delta_A + 2\mu &\rightarrow& {\rm hole~pocket}.
\end{eqnarray}
In our conventions a positive chemical potential favors hole pocket
formation
while disfavoring electron pockets. As $t^{eff}_i(k_y)$ grows beyond
this
minimal value, the pockets grow in size. We take the hopping such
that
\begin{equation}\label{eIIIviii}
t^{eff}_i (k_y-K_y) =
\begin{cases}
-t_0(1- (k_y-K_y)^2/\kappa_0^2 + \ldots ), &K_y\sim 0, \pm\pi/2 \cr
t_0(1- (k_y-K_y)^2/\kappa_0^2 + \ldots ), &K_y\sim \pm\pi/4,\pm 3\pi/4
\end{cases}
\end{equation}
where $\kappa_0$ is the small parameter guaranteeing that the RPA
is a good approximation.
The dispersion relations of the quasi-particles near the hole pockets are
\begin{equation}\label{eIIIix}
E_i(k_x,k_y) = \frac{(k_x-p_{i})^2}{2m_{||i}} + \frac{(k_y-K_y)^2}{2m_{\perp i}} - \epsilon_{Fi}
\end{equation}
where $p_{i} = \pm K_{FA_i} \mp \frac{t^{eff}_i(0)}{2v_{Fi}}$,
$\epsilon_{Fi} = \frac{(\gamma_it^{eff}_i(0))^2}{8m_{||i}v^2_{Fi}}$,
$\gamma_i = (1-\frac{4}{(t^{eff}_i(0))^2}(\Delta^2-\mu^2+\mu t^{eff}_i(0)))^{1/2}$,
$m_{\perp i} = \frac{\kappa^2_0}{2t^{eff}_i(0)}$, and $m_{||i}=\frac{t^{eff}_i(0)-2\mu}{2v^2_{Fi}}$.
In Figure 3 are plotted the expected Fermi pockets. On the l.h.s. of
Figure 3 are plotted the pockets found at zero chemical potential
while
on the r.h.s. are plotted the pockets for a chemical potential such
that
$2\Delta_A+2\mu> t_0 > 2\Delta_A-2\mu$. For such a condition
one obtains only hole pockets. We see that hole pockets occur
in the vicinity of $(\pm\pi/2,\pm\pi/2)$.
\subsubsection{Luttinger sum rule}
The Luttinger sum rule (LSR) for the single particle Green's functions
at the particle hole symmetric point takes the form
\begin{eqnarray}\label{eIIIx}
n = \frac{2}{(2\pi)^2}\int_{G(\omega = 0,k) >0}d^dk,
\end{eqnarray}
where $n$ is the electron density.
The corresponding Luttinger surface of $G(\omega,k)$ is defined as
the loci of points in $k$-space where $G(\omega=0,k)$ changes
sign. These sign changes occur both at the poles and the
zeros of $G$.
In order to apply the Luttinger sum rule, we must take $G(\omega,k)$
to be one of $G^{2D}_{A_{1/2}}$, i.e. we must apply the LSR to each
band separately (see Eqn. (\ref{eIIIiii}) for the definition of $G^{2D}_{A_{1/2}}$).
(We only apply the LSR to the electrons in the A-bands -- the LSR also
holds separately for electrons in the B-bands.)
At the particle-hole symmetric point, zeros are present in $G^{2D}_{A_{1/2}}(0,{\bf k})$ along
the lines $k_y=\pm K_{F_{A_i}}$. In the absence of pockets the LSR is satisfied
$G(\omega=0,k_x=K_{F_{A_{1,2}}},k_y)$ because of these zeros. And when $t^{eff}_i$ becomes
strong enough so that pockets form, the appearance of equally size electron
and hole pockets on either side of $\pm K_{FA_i}$ ensure that the Luttinger sum
rule continues to hold.
Introducing a finite chemical potential (with $\mu < \Delta_A/2$)
leaves the LSR violated as expressed in Eqn. (\ref{eIIIx}). However it continues to hold in a modified form.
Because in a finite chemical potential, the ladder Greens functions are given
by $G_{A_i}(\omega,\mu,k) = G_{A_i}(\omega-\mu,0,k)$, the LSR holds if
we consider the sign changes the Green's function undergoes not at $\omega=0$
but at $\omega=\mu$.
\subsubsection{Superconducting Instability}
The residual interactions between the Fermi pockets and the Cooperons
will lead to
instabilities in the RPA solutions as temperature goes to zero. Provided a finite chemical
potential is present the leading
instability will be to a superconducting state. While gapless
quasi-particles only exist in the A-bands, both A and B bands
will go superconducting simultaneously.
The general form of the Cooperon-quasiparticle interaction is
\begin{eqnarray}\label{eIIIxi}
H_{\phi QP} &=& \sum _{i=A, B;{\bf k},{\bf q} }
\frac{\Gamma_i({\bf k},{\bf q})}{(NLa)^{1/2}}
\Big[\Phi_i({\bf q})\Delta_{QPA}^\dagger({\bf k},{\bf q}) + h.c.\Big] \cr\cr
&+& \frac{1}{2}\sum_{{\bf q},{\bf k},{\bf k}'} \frac{g({\bf q},{\bf k},{\bf k}')}{NLa}
\Delta_{QPA}^\dagger({\bf k},{\bf q}) \Delta_{QPA}({\bf k}',{\bf q}) \cr\cr
\Delta_{QPA}^\dagger({\bf k},{\bf q}) &=&
\epsilon_{\sigma\s'} [A^\dagger_{1\sigma}({\bf k} +{\bf q})A^\dagger_{1\sigma'}(-{\bf k}) -
A^\dagger_{2\sigma}({\bf k} +{\bf q})A^\dagger_{2\sigma'}(-{\bf k}) ] .
\end{eqnarray}
Here $L$ is the length of the ladders, $a$ is the interladder spacing, and $N$ is the
number of ladders in the array.
$\Phi_{A,B}$ are the Cooperon fields whose bare propagators are defined as
\begin{equation}\label{eIIIxii}
D^0_{i}(\omega_n,k) = \langle T \Phi_i({\bf k},\omega_n) \Phi_i^\dagger ({\bf k},\omega_n)\rangle_0 =
\frac{v_{Fi}}{-(i\omega_n-2\mu)^2 + \Delta_i^2 +(v_{Fi}k_x)^2}.
\end{equation}
We see
that $g$ has the dimensionality of energy$\times$length$^{2}$
and $\Gamma_i$ has the dimensionality of energy$\times$length$^{1/2}$.
The different terms in Eqn. (\ref{eIIIxi}) have different origins.
The strongest interactions are presumably $\Gamma_i$ as this term
already exists for uncoupled ladders. Inter-ladder interactions, such as interladder
Coulomb repulsion, also contribute to $\Gamma_i$. However interladder hoping does not -- this
contribution is suppressed due to a mismatch between the Fermi momenta of the $A_{i}$ and $B_{i}$
bands. The coupling $g$ is smaller than $\Gamma_i$:
it arises only in second order perturbation theory from intraladder
interactions and from presumed weak inter-ladder Coulomb interactions.
The pair susceptibility for the quasiparticles $\Delta_{QP A}$ in an RPA approximation is given by
\begin{eqnarray}\label{eIIIxiii}
\chi^{RPA}_{QPA}(\omega_n,{\bf q}) &=&
\frac{1}{LNa}\sum_{{\bf k_1},{\bf k_2}}\int^\beta_0 d\tau e^{i \omega_n\tau}
\langle T \Delta_{QPA}({\bf k_1},{\bf q},\tau)\Delta_{QPA}^\dagger ({\bf k_2},{\bf q},0)\rangle \cr\cr
&=& \frac{2C(\omega_n,{\bf q})}{1+g({\bf q})C(\omega_n,{\bf q})-2\sum_i\Gamma^2_i({\bf q})C(\omega_n,,{\bf q})
D^0_i(\omega_n,{\bf q})}.
\end{eqnarray}
We have assumed that the couplings $g({\bf q},{\bf k},{\bf k'})$ and $\Gamma_i({\bf k},{\bf q})$ are such
that we can ignore their dependence on ${\bf k}$ and ${\bf k'}$.
Here $C(\omega_n,{\bf q})$ is the Cooper bubble:
\begin{eqnarray}\label{eIIIxiv}
C(\omega_n,{\bf q}) = 2\int \frac{dk_xdk_y}{4\pi^2}
\bigg[ \frac{f(\epsilon_{A1}({\bf k}+{\bf q}))-f(-\epsilon_{A1}(-{\bf k}))}{i\omega_n - \epsilon_{A1}({\bf k}+{\bf q}) - \epsilon_{A1}(-{\bf k})}
+ (\epsilon_{A1}\leftrightarrow\epsilon_{A2})\bigg].
\end{eqnarray}
Here $\epsilon_{A_{1/2}}(k)$ are the bare dispersions of the $A_{1/2}$ quasi-particles. As $T\rightarrow 0$,
$C(\omega_n,q=0)$ develops a logarithmic divergence:
$C(\omega_n,q=0) \approx \sqrt{m_{||}m_\perp} \log (\frac{\epsilon_F+\mu}{T})$.
The pair susceptibility for the Cooperons fields has a similar RPA form:
\begin{eqnarray}\label{eIIIxv}
\chi^{RPA}_i(\omega_n,{\bf q}) =
\langle T \phi_i(q,\tau)\phi_i^\dagger (q,0)\rangle &=& D^0_i(\omega_n,k)
+ (D^0_i(\omega_n,k))^2\Gamma_i^2(q)\chi^{RPA}_{QPA}(\omega_n,{\bf q})\cr\cr
&&\hskip -1in = \frac{D^0_i(\omega_n,k) + g(q)C(\omega,q)-2D^0_{\tilde i}(\omega_n,{\bf q})C(\omega_n,q)\Gamma_{\tilde i}^2(q)}
{1+g({\bf q})C(\omega_n,{\bf q})-2\sum_i\Gamma^2_i({\bf q})C(\omega_n,,{\bf q})
D^0_i(\omega_n,{\bf q})}.
\end{eqnarray}
where $\tilde A = B,\tilde B = A$.
The superconducting instability occurs when the denominator in Eqns. (\ref{eIIIiii} and \ref{eIIIv})
vanishes at $\omega_n=0, q=0$, that is
\begin{eqnarray}\label{eIIIxvi}
C(0,0)\bigg(g(0) - 2\sum_i\Gamma^2_i(0)\frac{v_{Fi}}{\Delta_i^2 -4\mu^2}\bigg) - 1=0.
\end{eqnarray}
We note that this vanishing occurs simultaneously in all channels.
If $g>0$ (though interladder Coulomb repulsion is repulsive, the interactions between
quasi-particles on a given ladder is attractive leaving the sign of g indeterminate)
the instability occurs only when the chemical potential
approaches sufficiently close to $\Delta_A/2$ so that the resulting effective
interaction becomes attractive.
This chemical potential corresponds to minimal doping at which the superconductivity appears.
Taking $\Delta_A \ll \Delta_B$, the corresponding
transition temperature takes the
form
\begin{equation}\label{eIIIxvii}
T_c = {\rm max} \{T_{c1},T_{c2}\}; ~~~
T_{c} \approx \epsilon_{Fi}\exp \frac{1}{\sqrt{m_{||i}m_{\perp i}}}
\bigg [\frac{2\Gamma^2_A(0)v_{FA}}{(\Delta^2_A-4\mu^2)}-g(0)\bigg]^{-1},
\end{equation}
where $\epsilon_{Fi}=\frac{\gamma^2_i}{4}\frac{(t^{eff}_i(0))^2}{t^{eff}_i(0)-2\mu}$.
If we suppose that $\mu$ and $t^{eff}_i$ are such that we only have
hole pockets, the density of dopants is equal to
\begin{equation}\label{eIIIxviii}
x(\mu) = \frac{\kappa_0}{2^{7/2}\pi^2|t^{eff}_i(0)|^{1/2}}\frac{(2\Delta_i+|t^{eff}_i(0)|+2\mu) (-2\Delta_i+|t^{eff}_i(0)|+2\mu)}{\sqrt{|t^{eff}_i(0)|+2\mu}}
\end{equation}
If we denote the critical doping as $x_c(\mu=\Delta_A/2)$ where the
A-Cooperon becomes soft, we see that the transition temperature
behaves as $T_{ci} \sim \exp(-\alpha (x_c-x))$ as $x$ approaches
$x_c$, that is to say, the transition temperature has a strong
dependence on doping.
It should be emphasized that this critical doping $x_c$ as defined
above does not coincide with the optimal doping
as typically understood.
Optimal doping can be thought of as the doping level associated with a change in the Fermi surface topology.
However in this understanding our model always remains in the underdoped
regime since the quasiparticle Fermi surfaces
remain small as far as the interladder tunneling remains much smaller
than the gap of the outer (B) band pair.
\subsection{Scenario 2}
We now consider the second scenario where $\Delta_B \gg t_\perp \gg \Delta_A$.
Because $t_\perp$ is much larger than $\Delta_A$ but smaller
than $\Delta_B$, the effects
of the interactions are wiped out in the A-bands while preserved in
the B-bands. In particular a gapful Cooperon still exists on the B-bands
while the coupled A-bands appear as an anisotropic two dimensional
Fermi liquid.
\begin{figure}
\centering
\epsfig{file=2dfl.eps,width=0.4\linewidth,clip=}
\caption{The Fermi surface of the A-bands in a periodic zone scheme.}
\end{figure}
We can distinguish two parameter ranges in this scenario. At small dopings $\mu < \Delta_B/2$,
the B-Cooperons remain gapped.
The effective Hamiltonian for the two dimensional Fermi
liquid in the A-bands and the Cooperons in the B-bands appears as
\begin{eqnarray}\label{eIIIxix}
H^{2D} &=& \sum_{\bf k} \epsilon_1({\bf k})A^\dagger_1({\bf k})A_1({\bf k}) +
\epsilon_2({\bf k})A^\dagger_2({\bf k})A_2({\bf k}) + \sum_{\bf
k}E_{B_c}(k_x)\Phi_B^\dagger({\bf k})\Phi_B({\bf k});\cr\cr
\epsilon_i({\bf k}) &=& E_{A_i}(k_x)+t^{eff}_i(k_y)\cr\cr
E_{B_c}(k_x) &=& \sqrt{k_x^2+\Delta^2_B}-2\mu,
\end{eqnarray}
where $E_{A_i}(k_x)$ is given in Eqn. (\ref{eIIi}) and
$t^{eff}_i(k_y)$ in Eqn. (\ref{eIIIiv}).
We illustrate the two dimensional Fermi surface of the A-bands in Figure
4.
The form of the quasi-particle-Cooperon interaction is that of
Eqn. (\ref{eIIIxiii}) (though of course, now we have no A-Cooperon and so this coupling
is absent).
This system, like in Scenario 1, has a pairing instability to superconductivity.
The pairing susceptibilities in an RPA approximation take a similar form as for Scenario 1:
\begin{eqnarray}\label{eIIIxx}
\chi^{RPA}_{QPA}(\omega_n,{\bf q}) &=&
\frac{2C(\omega_n,{\bf q})}{1+g({\bf q})C(\omega_n,{\bf q})-2\Gamma^2_B({\bf q})C(\omega_n,,{\bf q})
D^0_B(\omega_n,{\bf q})};\cr\cr
\chi^{RPA}_B(\omega_n,{\bf q}) &=&
\frac{D^0_i(\omega_n,k) + g(q)C(\omega,q)}
{1+g({\bf q})C(\omega_n,{\bf q})-2\Gamma^2_B({\bf q})C(\omega_n,,{\bf q})
D^0_B(\omega_n,{\bf q})}.
\end{eqnarray}
where $C(\omega_n,q)$ is defined as in Eqn. (\ref{eIIIxv}).
As we no longer have pockets as in Scenario 1, but instead have an anisotropic 2D Fermi liquid
whose Fermi surface consists of slightly deformed lines (see Figure 4), the divergent with temperature behaviour of
$C(0,0)$ now takes the form
\begin{equation}\label{eIIIxxi}
C(0,0) = \frac{1}{a\pi v_{FA}}\log(\frac{E_{FA_1}E_{FA_2}}{T^2}).
\end{equation}
Because the A-quasi-particles
are already gapless, a finite $\mu$ dopes the A-bands with doping $x^A(\mu)$.
Thus $E_{FA_i}(\mu) = E_{FA_i}(\mu=0)-\mu v_{FA}$. If we
denote the critical doping, $\mu_c$, as the doping when the B-Cooperon
becomes soft (i.e. $\mu_c=\Delta_B/2$) and $x^A_c = x^A(\mu_c)$ the
corresponding doping of the A-bands, we can rewrite the form
of the B-Cooperon propagator, $D^0_{B}(0,0)$, as
\begin{equation}\label{eIIIxxii}
D^0_{B}(\omega_n=0,q=0) \sim \frac{v_{FB}}{v^2_{FA}a^2((x^A_c)^2-x^2)}.
\end{equation}
Again we emphasize that the critical doping $x^A_c$ as defined
above does not coincide with optimal doping -- in this model we are
always
in the underdoped regime.
For this range of doping we obtain a transition temperature of the form
\begin{equation}\label{eIIIxxiii}
T_c = (\epsilon_{FA1} \epsilon_{FA2})^{1/2}
\exp\bigg[-\pi a v_{FA}\bigg(\frac{2\Gamma^2_B(0)v_{FB}}{v^2_{FA}a^2((x^A)^2-x^2)}-g(0)\bigg)^{-1}\bigg],
\end{equation}
and we see that the critical temperature grows extremely fast with
doping, similar to the transition temperature determined in Scenario I.
The second region occurs at $x > x_c$, when the holes penetrate into the outer B-bands. Here
the O(8) Gross-Neveu model governing the B-bands undergoes a crossover into a $O(6)\times U(1)$ Gross-Neveu model.
The B-Cooperon propagator at $\omega,k=0$ becomes more singular. At the same time the
velocity of the phase fluctuations becomes small and these fluctuations can be treated as
slow modes. Integrating over the nodal fermions one obtains the effective Lagrangian for the phase fluctuations:
\begin{eqnarray}\label{eIIIxxiv}
{\cal L} &=& \sum_{n}\Big[- J_c\cos\bigg(\frac{1}{2}(\phi_n(x) - \phi_{n+1}(x))\bigg) \cr\cr
&& + \frac{K(\mu)}{8\pi}\bigg(v_F(\mu )(\partial_x\theta_n-4\mu)^2
+v_F(\mu )(\partial_{\tau}\theta_n)^2\bigg) -
\frac{M}{2}\cos(\theta ),
\end{eqnarray}
where $n$ is a sum over ladders.
As we have already noted the parameter $K$ is renormalized by the
Coulomb interaction to be slightly less than 1. $v_F(\mu )$ is more
dramatically affected, taking the form $v_F(\mu ) \sim
v_{FB}(\frac{2\mu}{\Delta_B}-1)^{1/2}$
so that it vanishes at $x=x_c$ (or equivalently $\mu=\Delta_B/2)$.
As a side remark we note that there is an alternative way of presenting the effective Hamiltonian.
The above Lagrangian (Eqn. \ref{eIIIxxiv}) is the continuum limit of the following model:
\begin{eqnarray}\label{eIIIxxv}
H &=& \sum_{n,m}\Big\{ - J(\tau_{n,m+1}^+\tau_{n,m}^- + h.c.)
- J_c(\tau_{n+1,m}^+\tau_{n,m}^- + h.c.) + \cr\cr
&& \hskip .5in [(-1)^nM -2\mu ]\tau^3_{n,m}\Big\},
\end{eqnarray}
where $\tau^a$ are Pauli matrix operators. In the continuum limit $\tau^-$ becomes the order parameter
field $\mbox{e}^{i\frac{\phi}{2}}$. Here $J \sim M$.
The model presented above is a model of anisotropic spin-1/2 magnet on a 2D lattice with a staggered ($M$) and
uniform magnetic fields ($2\mu$). This form of the Hamiltonian has been proven to be very
convenient for numerical calculations yielding promising results for the transport.\cite{assa}
We again estimate the transition temperature using an RPA argument.
At $T=0$ the doping of the entire system (both the A and the B bands)
is
\begin{equation}\label{eIIIxxvi}
x = \mu\rho_A + c\frac{\Delta_B}{v_{FB}a}(\frac{2\mu}{\Delta_B}-1)^{1/2}
\end{equation}
where $c$ is a constant and $\rho_A=\frac{2}{av_{FA}\pi}$
The detailed form of the Cooperon propagator for a single chain at T=0
can be extracted from Ref \onlinecite{caux}.
However to obtain an estimate for $T_c$, it is enough to use the finite
temperature Luttinger liquid expression for the Cooperon propagator:
\begin{equation}\label{eIIIxxvii}
D^0_{B}(\tau,x) =
Z\bigg(\frac{1}{\epsilon_{FB}(\mu)\beta}\bigg)^{1/2K}\frac{1}{\sinh(T\pi(x/v_{FB}(\mu
) + i\tau))},
\end{equation}
where $Z$ is a numerical constant.
Thus
\begin{equation}\label{eIIIxxviii}
D^0_{B}(\omega=0,{\bf q}=0) \sim \frac{v_{FB}}{NL}(\frac{2\mu}{\Delta_B}-1)^{1/2-1/2K}T^{-2+1/2K}.
\end{equation}
Substituting the latter expression into RPA expressions for the
pairing susceptibilities (Eqns. \ref{eIIIxx}) we obtain
an estimate for the critical temperature upon doping as follows:
\begin{equation}\label{eIIIxxix}
T_c \sim \Big(x-\frac{\Delta_B\rho_A}{2}\Big)^{\frac{2-2K}{4K-1}}.
\end{equation}
This dependence on the doping is much weaker than (Eqn. \ref{eIIIxxvii}).
It holds in the region where phase fluctuations are already strong.
Thus we have obtained two regimes with different doping dependence of $T_c$.
The first one is the BCS-like with $T_c$ given by Eqn. (\ref{eIIIxxiii}).
It corresponds to lowest doping levels. The other regime, which in our model still
describes a situation an anisotropic 1D-like Fermi surface, is the regime with strong phase fluctuations.
The mean field transition temperature in this regime is given by Eqn. (\ref{eIIIxxix}). A further
increase of doping presumably will lead to a change in the Fermi surface topology and is
not considered in this paper.
\section{Discussion}
Phenomenological models based on coupled fermions and bosons similar to that derived here, have
been proposed much earlier Refs.(\onlinecite{lee,ranninger,gesh,chub}) to describe the high
temperature superconductors. The closest similarity are to the models proposed by Geshkenbein, Ioffe
and Larkin\cite{gesh} and by Chubukov and Tsvelik.\cite{chub,chub1} Both these phenomenological
models examined Fermi arcs centered on the nodal directions, coupled in the d-wave channel to Cooperons
associated with the antinodal regions. The model studied in Ref. (\onlinecite{gesh}) had dispersionless Cooperons which
provided BCS-style coupling for the nodal quasiparticles. The result was a superconducting transition with
weak fluctuations, similar to our $x < x_c$ case. In the model considered in Ref. \onlinecite{chub} the Cooperons
possessed a
one-dimensional dispersion which resulted in strong fluctuations as takes place in our case for $x > x_c$.
The authors of Refs. \onlinecite{gesh,chub,chub1} considered the fluctuation regime above T$_c$ when the
Cooperon energy is close to the chemical potential and drew comparisons to experiments in several underdoped cuprates.
The key ingredients controlling superconductivity in the array of 4-leg Hubbard ladders that we have considered
in this letter, are a small residual Fermi surface (either pockets as in Scenario I or arcs as in Scenario II),
which is coupled in the d-wave Cooper channel to a finite
energy Cooperon associated with the pseudogap responsible for the partial truncation of the Fermi surface.
The properties of a weak coupling 4-leg Hubbard ladder near to half-filling are used to obtain these key
ingredients. Our goal is to derive a tractable model containing the important features that are relevant
to high temperature superconductivity in the cuprates. In order to assess the relevance of our model to
this goal, clearly one must examine whether these key ingredients are present in a two dimensional
Hubbard model on a square lattice near half-filling.
As we mentioned above, earlier numerical renormalization group studies on the 2-dimensional Hubbard model
were interpreted as pointing towards a similar pairing mechanism arising from enhanced pairing correlations
present in a condensate that truncates the Fermi surface in the antinodal regions. There are of course two
reservations in these earlier works. Firstly, the one loop approximation in the numerical renormalization
group studies limits them to at most moderately strong onsite repulsive interactions. Secondly, the
renormalization group studies per se break down when the scattering vertices flow to strong coupling
and the nature of the resulting low energy or low temperature effective action is a difficult problem
which could only be surmised rather than explicitly derived. These two weaknesses make it imperative
to examine the question whether these key ingredients are present also for strong coupling.
The most reliable strong coupling calculations are exact diagonalization studies of strong coupling
Hamiltonians. The only limitation is the finite cluster size which currently is limited to small
clusters containing up to 32 sites and 1,2 and 4 holes. Leung and his collaborators
Refs. \onlinecite{leung1,leung2,leung3} have reported a series of calculations for
these clusters using the strong coupling t-J model and its extensions to include
longer range hopping and interactions. We begin with a recap of the main conclusions of these calculations.
The allowed set of {\bf k}-points in a 32-site cluster with periodic boundary conditions
contain both the four nodal ($\pi/2, \pi/2$) and two antinodal points ($\pi,0$) and ($0, \pi$).
A single hole enters at a nodal point. For two holes there are two different states that are
possible groundstates depending on the parameter values. For the plain t-J model with only
nearest neighbour hopping a 2-hole bound pair state with d($x^2 - y^2$) symmetry is the
groundstate on the 32-site cluster for J/t $>$ 0.28. The binding energy is
quite small at J/t = 0.3 but grows with increasing J/t. An extrapolation from finite size
clusters to the infinite lattice however suggests that the pair state is no longer the
groundstate at J/t =0.3, but an excited state with an energy of approximately 0.17t.\cite{leung1}
The inclusion of longer range interactions and hopping in the t-J model increases the energy of the
pair state further and confirms the conclusion that for parameter values relevant to cuprates the
groundstate of the cluster has two unbound holes in the nodal states.\cite{leung2} Extending the
calculations to the 32-site clusters with 4 holes, which corresponds to a doping of 1/8, shows all
4 holes entering into nodal states with no signs of pairing correlations.\cite{leung3} In view
of the prominent bound pair excited state for 2 holes, a low energy excited state with two of the
holes in a bound state may also be expected here. However at present there is no information on
this question to the best of our knowledge.
Leung and collaborators\cite{leung1,leung2,leung3} concluded from these calculations
that at low densities holes entered the nodal regions, possibly in pockets, and as a result there was
no evidence for d-wave pairing correlations in the groundstate for realistic values of the parameters
in t-J models. However the analysis presented here suggests a more optimistic conclusion. First we
note that the nodal points in the 32-site cluster are very special, because exactly at these points
the coupling in a Cooper channel to a d-wave Cooperon vanishes by symmetry. Thus if we interpret the
d-wave pair excited state as evidence for a finite energy Cooperon in the t-J model and its extensions,
then as the occupied
holes at finite doping move out from the exact nodal points, a d-wave pairing attraction is generated through
the coupling to this Cooperon, similar to the scenarios we discussed earlier. Note an earlier study for
two holes on smaller clusters by Poilblanc and collaborators\cite{Poilblanc} concluded in favor of the
interpretation of the 2-hole bound state as a
quasiparticle with charge 2e and spin 0, which would be an actual carrier of charge under an applied
electric field. In other words they concluded that a Cooperon is present in the strong coupling t-J model
at low doping. A more detailed analysis of the origin of the pairing in this state was published recently
by Maier et al.\cite{scal} Note the hole density in the case of 2 holes in a 32-site cluster is very
low so that the superconducting order we are postulating should coexist with long range antiferromagnetic
order. There is considerable evidence both numerical, in variational Monte Carlo calculations, and experimental,
in favor of such coexistence, as discussed in the recent review by Ogata and Fukuyama.\cite{ogfu}
We conclude that there is strong evidence that the pairing mechanism in the present model is not
confined to weak coupling and ladder lattices, but will also operate in the strong coupling t-J model
on a square lattice at low doping.
ATM and RMK acknowledge support by the US DOE under contract number DE-AC02-98 CH 10886.
TMR was supported by the Center for Emerging Superconductivity funded by the U.S. Department of Energy,
Office of Science and by MANEP network of Swiss National Funds.
|
\section{Introduction}
The Coma supercluster is the nearest rich supercluster of
galaxies \citep{cr76,gt78}, consisting of two rich Abell clusters,
separated by $30\,h^{-1}_{70}$~Mpc, but connected by a prominent
filament of galaxies and poorer groups \citep[{e.g.} ][]{font84}, which
is part of the supercluster identified as the ``Great Wall'' in the first major
redshift survey of galaxies in the nearby Universe \citep{gh89}. At a
distance of $\sim 100\,h^{-1}_{70}$~Mpc, it affords a closer look at
the properties of individual galaxies (1~kpc\,$\simeq$\,2.1 arcsec),
but its large angular scale on the sky presents observational
challenges for the narrow fields of view of most instruments.
It is interesting to note that even though they are the two most
significant structures in the supercluster, the two Abell clusters are
remarkably different in every respect. The optical luminosity
function of galaxies in the Coma cluster, for instance, has a much
steeper faint-end slope than that of its neighbour
Abell~1367 \citep{paramo03}. Early studies
\citep[e.g.][]{dress80} revealed that the fraction of spiral galaxies
in the Coma cluster is anomalously low compared to
Abell~1367. In addition, galaxies in the core of the Coma cluster were
identified to be unusually deficient in neutral hydrogen
\citep[e.g.][]{sj78,gh85,bern94}, ionized hydrogen \citep{sr97}
and molecular gas
(H$_2$/CO) \citep{boselli97a}.
\citet{bothun84} found a strong HI gradient in Coma galaxies-
many of the inner, HI poor spirals being quite blue, suggesting that some
gas removal process has acted quite recently, while Abell~1367 was found to
be a mixture of HI poor and rich galaxies.
The availability of more detailed analyses of star formation, based
on multi-wavelength data, and of spectral indices sensitive to
stellar ages, has made this field more interesting. Estimates of
star formation activity from mid-infrared (IR) \citep{boselli97b} and
Galaxy Evolution Explorer ({\em GALEX}) ultraviolet \citep{cortese08}
photometry show that star formation in Coma on the whole is
substantially suppressed compared to that in the field, while
Abell~1367 has an abundance of bright star-forming galaxies
\citep{cortese08}.
In a hierarchical model of the formation of structures in the
Universe, it is not surprising that a rich cluster, with a relaxed
appearance in its X-ray image, such as Coma, would represent the end
product of the gradual assimilation of several galaxy groups over
time. In spite of the claim of \citet{ds88} that the Coma cluster
does not have significant substructure, it has subsequently been shown
to have at least three major subclusters in the optical
\citep[{e.g.}][]{west98} and X-ray maps \citep[{e.g.}][]{dm93,adami05}. At
the other end of the supercluster, Abell~1367 is also found to be
elongated with three major subclusters, along the axis of the
filament, with a population of star-forming galaxies infalling into
the SE cloud and possibly the other two as well \citep{cortese04}.
This has prompted a multitude of studies attempting to link the
dynamical history of the supercluster with the properties of
individual galaxies to investigate both the process of building
clusters, as well as the effect of large-scale structure on the star
formation history of galaxies \citep[{e.g.}][]{chris06}.
By employing the spectral analysis of galaxies in Coma, together with
the X-ray map of \citet{neumann03}, \citet{poggianti04} found that the
post-starburst (k+A) galaxies, in which star formation has been
quenched within the last 1-1.5 Gyr, are associated with the X-ray
excess attributed to the substructure in Coma. However, this work was
limited to 3 fields in the $\sim\!2$~Mpc region surrounding the
centre of Coma.
In addition to star formation indices measured from optical spectra,
the availability of {\em Spitzer} MIPS mid-IR observations has been
recently utilised by several authors to characterise the star formation
properties of the obscured component in galaxies in clusters
and groups \citep[e.g.][]{saintonge08,chris09,wolf09,bai10}, since
the 24{$\mu$m}\ flux can be used as a representation of the
dust-reprocessed overall IR flux.
Hereafter, we will refer to the pair of clusters Coma and Abell~1367,
along with the associated filament of galaxies as the Coma
Supercluster. In this paper, we use the {\em Spitzer} MIPS 24{$\mu$m}\
observations, wherever available, along with the broadband colours
and spectral star formation indicators from the Sloan digital sky
survey (SDSS). We adopt the distance modulus of the Coma
Supercluster to be $m\!-\!M\!=\!35.0$, and use cosmological
parameters $\Omega_{\Lambda}\!=\!0.70, \Omega_{M}\!=\!0.30$, and
$H_0=70$ ~km~s$^{-1}$ Mpc$^{-1}$\ for calculating the magnitudes and distances. We
note that at the redshift of Coma (z\,$=\!0.023$), our results are
independent of the choice of cosmology. In the next section, we
present our data and reduction methodology, and summarise our main
results in \S\ref{coma-sf} and \S\ref{results}. We discuss their
implications, and compare the properties of galaxies in various parts
of the supercluster in \S\ref{discussion}, summarising in
\S\ref{conclusions}.
\section{Observational Data}
\label{data}
We base our work on the photometric and spectroscopic data acquired by the
SDSS Data Release~7 \citep[][(DR7)]{adelman06}, which for the first
time covers the entire region of the Coma Supercluster presented in
this paper. Over a smaller area, we also use archival 24{$\mu$m}~mid IR
(MIR) images, from the Multi-band Imaging Photometer (MIPS)
instrument aboard {\it{Spitzer}} \citep{rieke}, available over a significant
fraction of the two main clusters.
\subsection{Optical data}
We select the galaxies belonging to the Coma supercluster from the
SDSS spectroscopic catalogue only, requiring the member galaxies to
be within $170.0\!\le$RA$\le\!200.0$ deg and
$17.0\!\le$Dec$\le\!33.0$ deg on the sky, and with a radial
velocity within 2,000 km~s$^{-1}$ of the mean redshift of the Coma
cluster (6,973 km~s$^{-1}$) or the Abell 1367 cluster (6,495
km~s$^{-1}$) respectively \citep{rines03}. All our galaxies are
brighter than SDSS magnitude $r\!=\!17.77$ ($\sim\, M^{*}\!+\!4.7$ for
Coma), which is the completeness limit of the SDSS spectroscopic
galaxy catalogue.
\subsection{MIPS 24{$\mu$m}\ data}
For the mid-infrared study of Coma and Abell 1367, we use archival
24{$\mu$m}~{\em Spitzer} MIPS data covering $2{\times}2\,{\rm deg}^{2}$ in
the case of Coma (PID: 83, PI G.~Rieke) and
$30^\prime{\times}30^{\prime}$ for Abell 1367 (PID: 25, PI
G.~Fazio). The Coma 24{$\mu$m}\ dataset consists of four contiguous mosaics
(see Fig.~\ref{scl}) obtained in medium scan mode, with scan leg
spacing equal to the half array width, producing homogeneous
coverage over the mosaic, with an effective exposure time per pixel
of 88\,sec. \citet{bai}, who used this data to determine the
24{$\mu$m}~luminosity function of Coma, estimate the 80\% completeness limit
to be 0.33\,mJy, corresponding to a star formation rate (SFR) of
0.02\,M$_{\odot}\,{\rm yr}^{-1}$ at the redshift of Coma. The Abell
1367 24{$\mu$m}\ dataset consists of a single mosaic obtained in medium scan
mode, with scan leg spacing equal to the full array width, producing
an effective exposure time per pixel of 40\,sec.
The {\sc SExtractor} package \citep{bertin} was used to automatically detect
sources, and obtain photometric parameters. The images were first
background-subtracted, and then filtered with Gaussian functions,
with full width half maximum (FWHM) matching the 24{$\mu$m}\ point-spread
function (PSF). Aperture photometry was obtained for all objects
having 8 contiguous pixels above the 1$\sigma$ rms background noise
level. Following SWIRE, we measured the fluxes within circular
apertures of diameter 21, 30, 60, 90 and 120$^{\prime\prime}$. For
the vast majority of sources, the 6$^{\prime\prime}$ FWHM
point-spread function of MIPS leaves the object unresolved at 24{$\mu$m},
and hence we estimate the total 24{$\mu$m}\ flux of these objects from the
flux contained within the 21$^{\prime\prime}$ aperture, corrected by
a factor 1.29.
However, at the redshift of the Coma and Abell 1367 clusters, many of
the brighter galaxies (spirals in particular) are larger than our
nominal 21$^{\prime\prime}$ diameter aperture, such that a
significant fraction of their 24{$\mu$m}\ flux is missed. For these
galaxies, we use instead one of the larger apertures, matched to
contain the optical flux as parametrized by R$_{90z}$ (the radius
which contains 90\% of the $z-$\,band flux) as measured from the SDSS
image. Finally, for some of the bright cluster galaxies (9 in total),
the value of R$_{90z}$ is under-estimated. To correct for this, SM
eyeballed the 24{$\mu$m}\ images of these objects, and estimated an
appropriate radius, in each case, for extracting the `total' 24{$\mu$m}\
flux. In the case of the brighter spiral galaxies in Coma and Abell
1367, the 24{$\mu$m}\ emission often shows significant structure due to
spiral arms and star-forming regions, necessitating a high value of
the deblending parameter, to prevent the galaxy being shredded by
{\sc SExtractor}. Particular care was required to correctly deblend
the face-on spiral NGC\,3861 from NGC\,3861B, while keeping NGC\,3861
intact. We note that for $\sim$\,50\% of our sample the typical
measurement error in the 24{$\mu$m}\ flux is $<\!0.1$ mJy (which is
$\lesssim\!10\%$ of the measured flux), and for $>\!95\%$ the error
is $<\!0.35$ mJy. In total, $\sim$\,90\% of our galaxies have 24{$\mu$m}\
flux measured with $<\!30\%$ uncertainty.
\subsection{Matching the SDSS spectroscopic catalogue with the 24{$\mu$m}\
sources}
To match the SDSS spectroscopic catalogue for the Coma supercluster
with the MIR 24{$\mu$m}\ data, we choose the nearest optical counterpart
within 5$^{\prime\prime}$ of each 24{$\mu$m}\ source.
This accounts for a displacement of no more
than 2.5~$h^{-1}_{70}$ kpc between the centres. We note that
$\gtrsim$80\% of the matches are found to have centres within
2$^{\prime\prime}$ of each other. 16 of our sources above the
detection threshold were found to have multiple matches within
the matching radius of 5$^{\prime\prime}$.
Our final dataset comprises of 3,787 galaxies ($r\!\leq\!17.77$) with
spectroscopic redshifts from SDSS, and radial velocity within
$\pm$2,000 km~s$^{-1}$ of Coma or Abell 1367, in the $\sim$500 square
degree region of the sky covered by the Coma supercluster. Out of
these, 197 within $\sim$2 $h^{-1}_{70}$~Mpc of the centre of Coma and
24 within $\sim$0.65 $h^{-1}_{70}$~Mpc of the centre of Abell 1367
are found to have a 24{$\mu$m}\ counterpart. The probability that a background
MIPS source lies within the matching radius of 5$^{\prime\prime}$
by chance, depends only on the density of the 24{$\mu$m}\ sources. For the
completeness limit (0.33 mJy) adopted here, this density is
$\sim\!3,000$\,deg$^{-2}$ \citep[see fig. 3 of][]{shupe08}. This gives
the required probability as 0.018, implying that our sample of 197
24{$\mu$m}\ detected galaxies in Coma may have 3--4 interlopers.
A sample of the catalogue of data for the 24{$\mu$m}\ detected galaxies is
provided in Table~\ref{tbl:sb-table}.
\begin{table*}
\begin{minipage}{\linewidth}
\centering{
\caption{Catalogue of Coma galaxies detected at 24$\mu$m
(This table is available in full online.)}
\begin{tabular}{|c|c|c|c|c|c|c|c|}
\hline
\hline
RA & Dec & z & $g$ & $r$ & $z$ & 24$\mu$m flux & flag\footnote{
Classification on the basis of BPT diagram,where \\
0: unclassified \\
1: star-forming \\
2: AGN \\
3: AGN according to \citet{miller03} criteria (see text) \\} \\
(J2000) & (J2000) & & mag & mag & mag & (mJy) & \\
\hline
194.4574 & 28.6243 & 0.0250 & 17.80 & 17.35 & 17.24 & 0.31 & 1 \\
195.1548 & 28.6641 & 0.0236 & 17.47 & 16.89 & 16.49 & 1.25 & 1 \\
194.5385 & 28.7086 & 0.0255 & 15.21 & 14.56 & 14.08 & 36.02 & 1 \\
194.9172 & 28.6308 & 0.0179 & 15.67 & 15.47 & 15.30 & 7.90 & 1 \\
194.3645 & 28.4397 & 0.0260 & 17.16 & 16.57 & 16.19 & 0.75 & 1 \\
194.5628 & 28.5218 & 0.0231 & 17.00 & 16.56 & 16.32 & 1.84 & 1 \\
194.3912 & 28.4823 & 0.0209 & 14.35 & 13.54 & 13.00 & 1.86 & 3 \\
194.4747 & 28.4998 & 0.0244 & 15.94 & 15.18 & 14.56 & 0.04 & 0 \\
195.2170 & 28.3661 & 0.0255 & 14.51 & 13.73 & 13.10 & 1.39 & 0 \\
195.3123 & 28.5217 & 0.0281 & 17.48 & 17.11 & 16.98 & 1.06 & 1 \\
\hline
\label{tbl:sb-table}
\end{tabular}}
\end{minipage}
\end{table*}
\begin{figure*}
\centering{
{\rotatebox{270}{\epsfig{file=figure-1,width=13cm}}}}
\caption{{\bf{Top panel:}} The surface density of galaxies in the Coma
supercluster is shown in {\it grey}. The positions of passive
galaxies ({\it{black dots}}), AGN host galaxies ({\it{open red
circles}}) and starburst galaxies {\it{(blue stars)}};
EW(H$_{\alpha})\geq 25$\AA$\sim$ log SSFR$\sim -10$yr$^{-1}$) are
indicated. The {\it{big green circles}} are the groups in the
region, from the NASA Extragalactic Database. The rectangular pink
regions show the Spitzer MIPS fields of view for the two
clusters. {\bf{Bottom panel:}} The {\it{solid red}} line and the
{\it{dashed blue}} line show the fraction of AGN and of the
starburst galaxies, among all the galaxies shown in the upper
panel. The bins are chosen to have 150 galaxies in each of them.
The solid vertical lines represent the centres of the Coma cluster
and Abell~1367 respectively, while the {\it dot-dashed line} shows
the RA position of the NGC\,4839 group (see text). All the dotted
lines represent the positions of groups, shown
in the upper panel. }
\label{scl}
\end{figure*}
\section{Star formation and AGN across the Coma supercluster}
\label{coma-sf}
Located $\sim$100~Mpc from us, the Coma supercluster
offers a unique opportunity for investigating the effect on individual
galaxies of the hierarchical formation of structures in the Universe,
as galaxies progress through various environments towards the cores
of the clusters.
In order to give a general overview of the optical star formation
properties of galaxies along the entire supercluster, in
Fig.~\ref{scl} we plot the positions of all the galaxies from the
SDSS spectroscopic catalogue found in our redshift range, and the
major galaxy groups obtained from the NASA Extragalactic Database
(NED, http://nedwww.ipac.caltech.edu/). In order to avoid multiple
detections, we consider the groups only from the
NGS, WBL, USGC and HCG catalogues,
and even amongst these, we identify duplication by merging groups
closer than 1$^{\prime}$ and within $\pm$100 km~s$^{-1}$
of each other.
The positions of AGN host galaxies, identified using the BPT diagram
\citep[][also see \S\ref{sf}]{bpt81}, and non-AGN starburst galaxies
(star-forming on the BPT diagram, and having
EW(H$_\alpha$)$\geq$25\AA), are indicated in Fig.~\ref{scl}.
We also show the limited regions at the cores of the Coma and Abell 1367
clusters, over which the 24{$\mu$m}\ data are available, and utilised
in this paper.
Plotted in the bottom panel of the same figure, is the distribution
of the AGN and starburst galaxies, as fractions of the 150 galaxies
contributing to each bin. It is worth mentioning that although these
distributions are a fair representation of the data, the binning has
resulted in a spurious feature at the position of Abell 1367. As is
already well known in the literature, the central region of Abell 1367 is
sparsely populated, the majority of galaxies in the core being
late-type. In this figure, it appears that the fraction of
starburst galaxies does not decline in the core of Abell 1367, but if
the data is re-binned so as to have equally-spaced bins, this would not
be the case. Having said that, the decline in the fraction of
starburst galaxies at the centre of Abell~1367 is much shallower than
that seen in Coma.
\begin{figure*}
\centering{{\epsfig{file=figure-2,width=17cm}}}
\caption{The local AGN fraction ($f_{AGN}$) among massive galaxies
($z\!<\!14.5$), as a function of spatial position across the Coma
supercluster. The map is colour-coded with $f_{AGN}$.
The {\it{black contours}} indicate the
local luminosity-weighted ($z$-band) galaxy density across the
supercluster. The {\it{white contours}} indicate regions with AGN
fraction $f_{AGN}=$ 1 and 2-$\sigma$ below the mean
value over the supercluster. }
\label{agn_frac}
\end{figure*}
In general, the optically-identified AGN seem to be more or less
uniformly distributed throughout the supercluster (Figs.~\ref{scl}
and \ref{agn_frac}), except for a sharp decline in the cores of both
the clusters. This latter result is counter-intuitive, because AGN
hosts are known to be early-type massive galaxies which mostly reside
in dense environments. Hence, this may indicate that the optical
emission from the AGN present in this region is obscured and/or
detectable at other wavelengths (radio/X-ray).
The latter is beyond the scope of this work, but our
MIPS 24{$\mu$m}\ data supports the former possibility. As seen in
Fig.~\ref{radius}, not only does the relative fraction of the red
24{$\mu$m}\ galaxies marginally increase towards the centre, around 50\% of
them are found within 0.5~$h^{-1}_{70}$~Mpc of the centre of the Coma
cluster, indicating that optical emission from several AGN hosts may
be obscured. Further evidence in support of this argument may be
drawn from Fig.~\ref{bpt} where a large fraction of galaxies detected
at 24{$\mu$m}\ do not show [OIII] and/or H$_\beta$ in emission, but their
H$_\alpha$/[NII] flux ratios are consistent with the presence of an
AGN \citep{miller03} on the BPT diagram \citep{bpt81}.
We further investigate the environmental trends for AGN and star
formation activity within the Coma supercluster, by estimating the
local fraction of star-forming galaxies ($f_{SF}$) and AGN
($f_{AGN}$) over the entire supercluster, both for massive galaxies
\citep[$z\!<\!14.5$; $M_{z}{<}M^{*}+1.8$, assuming
$M_{z}^{*}{=}-22.32$ from][]{blanton01} and the dwarf galaxy
population ($z>15$; $M_{z}>M^{*}+2.3$). Following \citet{chris07}, we
define the local projected density of galaxies ($\rho(\mathbf{x})$)
using a variant of the adaptive kernel estimator \citep{silverman86,
pisani93}, where each galaxy $i$ is represented by an adaptive
Gaussian kernel $\kappa_{i}(\mathbf{x})$. This is different from the
algorithm of \citet{silverman86}, which, in its previous applications
to astronomical data, requires the kernel width $\sigma_{i}$ to be
iteratively set to be proportional to $\rho_{i}^{-1/2}$. In this
work, we fix the transverse width $\sigma_{i}$ to be proportional to
$D_3$, where $D_3$ is the distance to the third nearest neighbour
within 500\,km\,s$^{-1}$.
The dominant factors governing the star formation properties of a
galaxy are the mass of its DM halo, and whether it is the central
or a satellite galaxy in a halo \citep[{e.g.}][]{kauffmann04,
yang05, blanton06}. We adopt the above method, and choose the
dimensions of the kernel, keeping this in mind. In the case of
galaxies within groups or clusters, the local environment is measured
on the scale of their host halo (0.1--1 Mpc), while for galaxies in
field regions the local density is estimated by smoothing over their
5--10 nearest neighbours or on scales of 1--5\,Mpc \citep[for details, see][]{chris07}.
We can then map the local galaxy density ($\rho$) and $z$-band
luminosity density ($j_z$) as
\mbox{$\rho(\mathbf{x})=\sum_{i}\kappa_{i}(\mathbf {x{-}x}_{i})$} and
\mbox{$j_{z}(\mathbf{x})=\sum_{i}L_{z,i}\,\kappa_{i}(\mathbf
{x{-}x}_{i})$}, where $L_{z,i}$ is the $z$-band luminosity of
galaxy $i$ (see Fig.~\ref{sf_frac}). Following \citet{yang05}, we
identify local maxima in the $z$-band luminosity density as galaxy
groups and clusters, whose masses correlate with the total $z$-band
luminosity associated with the peak. By comparison with the
Millennium simulation, we expect all groups with 4 or more members to
be associated with a local maximum in \mbox{$j_{z}(\mathbf{x})$}
\citep{chris07}. The local $z$-band luminosity density is shown as
black contours in Figs.~\ref{agn_frac}--\ref{mean_ha}.
In Fig.~\ref{agn_frac} we show the spatial variation of $f_{AGN}$
among massive galaxies ($z\!<\!14.5$) across the supercluster, where
we calculate $f_{AGN}$ as
$\sum_{i{\subset}AGN}\kappa(\mathbf{x{-}x}_{i})/\rho(\mathbf{x})$,
where $i{\subset}AGN$ is the subset of galaxies classified as optical
AGN, based on the ratios of thermally excited and recombination lines
used in the BPT diagram. We do not consider the lower mass galaxies,
as the fraction of AGN declines rapidly to zero below
$M_z\!>\!M^{*}\!+\!2$;
For comparison, Fig.~\ref{sf_frac} shows the distribution of
$f_{SF}(\mathbf{x})$ as
$\sum_{i{\subset}SF}\kappa(\mathbf{x{-}x}_{i})/\rho(\mathbf{x})$,
where $i{\subset}SF$ is the subset of galaxies classified as
star-forming according to the line-ratios used in the BPT
diagram. Fig.~\ref{mean_ha} shows a map of the local mean equivalent
width (EW) of H$\alpha$ of star-forming galaxies as
$\langle{EW(H\alpha)}(\mathbf{x})\rangle=\sum_{i{\subset}SF}{\rm
EW(H}\alpha)\,\kappa(\mathbf{x{-}x}_{i})/\sum_{i{\subset}SF}\kappa(\mathbf{x{-}x}_{i})$,
plotted in a manner analogous to that used by \citet{chris06} for
galaxy colours in the Shapley supercluster.
\begin{figure*}
\centering{{\epsfig{file=figure-3a,width=17cm}}}
\centering{{\epsfig{file=figure-3b,width=17cm}}}
\caption{The local fraction of star-forming galaxies ($f_{SF}$) as a
function of spatial position across the Coma supercluster, for
massive galaxies ($z\!<\!14.5$; {\em top panel}) and dwarf galaxies
($z\!>\!15$; {\em lower panel}). The colours indicate $f_{SF}$, and
the same colour scale is used in
both the panels. Overlaid are black contours indicating the
$z$-band luminosity-weighted galaxy density across the
supercluster. This figure shows that in the Coma supercluster, star
formation in the massive galaxies ($z\!<\!14.5$) seems to be
suppressed independent of their local environment, while the dwarf
galaxies ($z\!>\!15$) are star-forming everywhere except in the
dense environment in the vicinity of rich clusters and
galaxy groups.
}
\label{sf_frac}
\end{figure*}
\begin{figure*}
\centering{{\epsfig{file=figure-4,width=15cm}}}
\caption{The local mean H$_\alpha$ equivalent width of star-forming
(EW(H$_\alpha)\!>\!2$\AA) dwarf ($z\!>\!15$) galaxies across the
Coma supercluster, indicated by colour. Overlaid are {\it{black contours}}
indicating the $z$-band luminosity-weighted galaxy density across
the supercluster, and {\it{white contours}} showing regions where
the local observed mean value of the EW(H$_\alpha$) is 2 and
3-$\sigma$ above that averaged over the whole supercluster. }
\label{mean_ha}
\end{figure*}
In Fig.~\ref{agn_frac} we see that the AGN fraction $f_{AGN}$,
among massive galaxies ($z\!<\!14.5$) declines at
the centre of the Coma and Abell 1367 clusters, while elsewhere, it is
uniformly distributed.
This appears to contradict the results of
\citet{chris07}, who found the AGN fraction to be independent of
environment, and to be a monotonically increasing function of stellar
mass. However, the volume studied by \citet{chris07} within SDSS DR4
did not cover clusters as rich as Coma or Abell 1367. This may
indicate that only the very dense environments affect the optical AGN
activity of galaxies. The relation between $f_{AGN}$ and local
environment is unclear in the intermediate density group
environment. While the $f_{AGN}$ values decline in the centre of some
groups, in others they exceed the mean value for the field. Although
there is a slight indication that the groups in which the $f_{AGN}$
appears to be declining, mostly lie on the filament connecting the Coma
and Abell 1367, or in the vicinity of the clusters themselves, while
the groups with higher values of $f_{AGN}$ lie in underdense regions.
However, it is not possible to draw any firm statistical inferences
from this observation.
Whether the AGN fraction $f_{AGN}$ varies with environment or not
depends on how the activity of the galactic nuclei is defined.
\citet{miller03} do not find any correlation between galaxy density
and f$_{AGN}$ in a large sample of galaxies ($M_r\!=\!-20;
0.05\!<\!\rm z\!<\!0.095$), where the AGN are defined in terms of
optical emission line ratios characterised by the BPT diagram. Other
studies find higher incidence of optical AGN in groups and clusters
\citep{arnold09}. If the AGN were selected according to their radio,
mid-IR or X-ray properties, they would be found in differen hosts:
radio AGN in early-type galaxies, IR AGN in bluer galaxies, and X-ray
AGN in galaxies of intermediate colour \citep{hickox09}. The
environment dependence of AGN activity is thus largely linked to the
distribution of the hosts. The incidence of X-ray AGN is higher in
galaxy groups than in galaxy clusters \citep{martini06,shen07}. The
fraction of radio-loud AGN is the same in the brightest galaxies of
groups and clusters and in the field, but higher in non-central
galaxies \citep{best05}. \citet{wdp10} show that the mass function
of black holes is independent of environment, and the variation in
the distribution of optical, radio and X-ray AGN can be understood in
terms of the accretion processes that lead to the manifestation of
the AGN in the various ranges of electromagnetic radiation.
We measured the significance of the spatial variations seen in
$f_{AGN}$ across the supercluster by performing Monte Carlo
simulations, in which we made repeated maps of $f_{AGN}$ after
randomly assigning the positions of the AGN to the bright galaxies
({i.e.}\ testing the null hypothesis that $f_{AGN}$ is constant across
the supercluster), and measuring the fraction of maps in which a
given local value of $f_{AGN}$ was obtained within one of the Monte
Carlo simulations. The results of these simulations are represented
by the white contours, which indicate regions that have $f_{AGN}=$
1 and 2-$\sigma$ below the mean value across the
supercluster. This confirms that the decline in $f_{AGN}$ seen
towards the core of the Coma cluster is significant at the 3$\sigma$
level, while that seen in Abell 1367 is significant at the 2$\sigma$
level.
In Fig.~\ref{sf_frac} we show the general correlation between the
fraction of star-forming galaxies ($f_{SF}$) and environment for
massive galaxies ($z\!<\!14.5$; top panel) and the dwarf galaxies
($z\!>\!15$; bottom panel). For the massive galaxies we find an
almost uniform $f_{SF}\!<\!0.5$ in all the environments. For dwarf
galaxies, we see a much stronger SF-density relation, with $f_{SF}$
rising rapidly from the $f_{SF}\!<\!0$.1--0.4 seen in the cluster
cores, to $f_{SF}\!>\!0.95$ in the field. This result is in agreement
with literature, where, using colour or SFR indicators,
it has been shown that dwarf galaxies exhibit much stronger
radial trends with environment than their massive counterparts
\citep[][among others]{gray04,tanaka,smith06,chris07}.
The origin of the SF-density relation for the
dwarf galaxies can be attributed to the fact that (i) the star
formation in dwarf galaxies can easily be quenched by the tidal
impact of a massive neighbour and/or the ICM of the cluster
\citep[{e.g.}][]{larson}, and (ii) unlike their massive counterparts,
dwarfs (of the luminosities considered here) do not become passive
by internal mechanisms such as gas consumption
through star formation, merging etc.. However, the effects of supernovae
wind blowouts and feedback become significant for relatively faint
($M^*\!\lesssim\!10^7$ M$_\odot$) dwarfs \citep[{e.g.}][]{maclow99,marcolini06}.
In this work we do not find any evidence for the quenching of star
formation in galaxies on the filament(s), but only in the cores of
the embedded galaxy groups.
The H$_\alpha$ emission traces the current SFR of a galaxy, while the
continuum flux under this line can be used as an indicator of its
past SFR, making the globally averaged EW(H$_\alpha$) from a galaxy
an effective indicator of the birthrate parameter ($b\!\equiv$
current SFR normalized by the SFR averaged over the lifetime of a
galaxy). Recently, \citet{lee09} have shown that
EW(H$_\alpha$)\,$\sim\!40$\AA~corresponds to $b\!=\!1$. In
Fig.~\ref{mean_ha} we show the variation in the mean EW(H$_\alpha$)
as a function of environment for the star-forming dwarf galaxies.
This is potentially a very powerful technique for dissociating the
intrinsically active star formation history (SFH) of low-mass
galaxies from a starburst caused by the impact of local environment
\citep[see][for instance]{chris07}.
If galaxies are slowly quenched by
interactions with their environment \citep{balogh04a}, a decline in
the mean EW(H$_\alpha$) is expected in the regions of denser
environment. On the other hand, if star formation is
triggered due to the impact of environment, the
mean EW(H$_\alpha$) should increase. Unlike the slow quenching of
dwarfs in the dense environments \citep{tanaka,chris07}, here, in the
Coma supercluster, we find that the dwarf galaxies follow different
evolutionary paths in the groups and in the denser cluster
environments. In galaxy groups, a dwarf galaxy is slowly quenched via
interactions with the other group members and/or the tidal field of
the group, while in the clusters, an infalling dwarf experiences a
starburst in the intermediate density environment at the cluster
periphery \citep{porter08,mrp10}, and is then rapidly
quenched via cluster-related environmental mechanisms, such as
ram-pressure stripping.
In analogy to Fig.~\ref{agn_frac}, we measure the significance of the
spatial variations seen in the mean EW(H$_\alpha$) via Monte Carlo
simulations, this time by making repeated maps after randomly
swapping the values of EW(H$_\alpha$) among the dwarf star-forming
galaxies. In doing this, we seek to test the null hypothesis that the
EW(H$_\alpha$) is independent of spatial position among star-forming
galaxies. In Fig.~\ref{mean_ha}, we show, as overlaid white contours,
the regions in which the local observed mean value of the
EW(H$_\alpha$) is 2 and 3$\sigma$ above that averaged over the whole
supercluster. This confirms that the excess star formation seen in
the infall regions of Coma, and towards the core of Abell~1367, is
indeed significant at $>$\,3$\sigma$ level. In quantitative terms,
this excess in the infall regions of the Coma cluster is due to a
population of $\sim$\,30 dwarf starburst galaxies ($z\!>\!15$,
EW(H$_\alpha)\!>\!40$\AA) located within or along the white
2-$\sigma$ contours in Fig.~\ref{mean_ha}. The analogous excess in
the core of Abell~1367 can be ascribed to $\sim$\,10 starburst dwarf
galaxies.
\section{Optical and mid-IR analysis of Coma and Abell~1367 galaxies}
\label{results}
The infrared (IR) emission from normal galaxies around
$\lambda{=}24\mu$m is dominated by hot dust in the H{\sc ii} regions
of massive stars, in addition to (usually) minor contribution from
asymptotic giant branch (AGB) stars and the general interstellar
medium (ISM). This recycled emission, together with the optical
emission, can thus provide a good estimate of the total starlight of
a galaxy and be used to better constrain the current and past star
formation rate (SFR) of a galaxy \citep[][among
others]{calzetti,kennicutt09,rieke09}. While IR astronomers usually
study late-type galaxies, early-type galaxies have received sporadic
attention only in the context of the warm dust component detected in
a few of them \citep[{e.g.}][]{knapp}.
In this paper, we combine the optical photometric and spectroscopic
data taken by the SDSS DR7, with the 24{$\mu$m}\ {\it{Spitzer}}/MIPS data,
which is sensitive to processed optical emission from stars, to
study the star formation activity of galaxies residing in the denser
regions of the Coma supercluster.
\subsection{Optical and MIR colours}
\label{colours}
\begin{figure}
\centering{
{\rotatebox{0}{\epsfig{file=figure-5,width=10cm}}}}
\caption{The $(g\!-\!r)$ vs $r$ colour-magnitude relation for the
24{$\mu$m}\ bright galaxies detected in the Coma cluster ({\it{circles}})
and Abell~1367 ({\it{big stars}}). The solid line shows the
colour-magnitude relation (CMR) fitted to the Coma data points,
while the dot-dashed lines mark the mean absolute deviation (MAD)
boundaries on either side. We adopt the lower MAD boundary in
$(g\!-\!r)$ for segregating the red sequence from the blue
cloud. Note that we do not use any upper bound for the red sequence
because fitting is done using spectroscopically confirmed cluster
members only. The symbols are colour-coded according to the
$(24\!-\!z)$ colour of the MIR bright sources (Fig.~\ref{ssf}). All
the other Coma galaxies are shown as {\it{grey stars}}. Note that
$(24\!-\!z)$\,=\,-6 mag separates the red sequence galaxies from
the blue ones in a $z\!-\!(24\!-\!z)$ plane (Fig.~\ref{ssf}). }
\label{cmr}
\end{figure}
\begin{figure}
\centering{
{\rotatebox{270}{\epsfig{file=figure-6,width=7cm}}}}
\caption{The $(24\!-\!z)$ colour of all galaxies in the Coma
supercluster sample, divided into blue cloud ({\it blue stars})
and red sequence ({\it red}) (Fig.~\ref{cmr}) galaxies, are plotted
against their $z$-band magnitudes. The dotted line shows the 80\%
completeness limit of the $24\mu$m data. The constant SFR lines
are drawn using the conversion factor given by
\citet{calzetti}. The horizontal dashed line at $(24\!-\! z)\!=\!-6$
mag is our empirically chosen criterion for selecting star-forming
galaxies. The top axis shows the stellar mass of galaxies estimated
using a relation from \citet{bell03}. }
\label{ssf}
\end{figure}
In Fig.~\ref{cmr}, we show the colour magnitude relation (CMR) for
all the spectroscopic galaxy members ($r\!\leq\!17.77$) found in this
region in the SDSS DR7, and those that are also detected at 24{$\mu$m}\
in the {\em Spitzer}/MIPS observations. For
comparison, the 24{$\mu$m}\ galaxies found in Abell~1367 are also
shown. We only use galaxies brighter than $r\!=\!15.5$ to fit the CMR.
The CMR is of the form $g\!-\!r\!=\!1.244\!-\!0.032\,r$ for
all the Coma galaxies, where the mean absolute
deviation from the relation is
$\!\pm\!0.064$ mag.
We repeat this analysis in Fig.~\ref{ssf} for the near-infrared (NIR)
band of SDSS ($z$-band), and the 24{$\mu$m}\ MIPS band. The galaxies
classified as red and blue on the optical colour-magnitude diagram,
split into two separate classes around $(24\!-\!z)= -6$ mag on the
plot of the near/mid IR colour and magnitude as well. Although we
overplot the lines of constant SFR in Fig.~\ref{ssf}, according to the
empirical relation given by \citet{calzetti}, it is important for the
reader to consider that the 24{$\mu$m}\ flux alone is an accurate SFR
indicator only for the late-type, dusty star-forming galaxies (we
return to this issue below). The SDSS $z$-band, centred
at $\sim$\,9000\AA, is a good measure of the light from evolved
stars, and hence the stellar mass of a galaxy. The top axis shows the
stellar mass of galaxies estimated using the relation $\log
M^*\!=\!-0.306\!+\!1.097(g-r) \!-\!0.1\!-\!0.4(M_r\!-\!5\log\,h
\!-\!4.64)$, from \citet{bell03}. On the other hand, the 24{$\mu$m}\ MIPS band in
the MIR is a good proxy for the dust-processed light. This makes the
$(24\!-\!z)$ colour an excellent approximation for the value of the
specific star formation rate (SSFR or SFR/M$^*$) of galaxies.
Figs.~\ref{cmr} and \ref{ssf} together show that the 24{$\mu$m}\ detected
red sequence galaxies in Coma have consistent optical and MIR
colours.
We note that the optical-MIR colours of the red sequence galaxies are
not consistent with those expected from photospheric emission from
old stellar populations, with an excess emission always apparent at
24{$\mu$m}. The {\em Spitzer} Infrared Spectrograph observations of
early-type galaxies in Virgo and Coma clusters show that the diffuse,
excess emission, apparent at 10--30{$\mu$m}\ in these galaxies, is due to
silicate emission from the dusty circumstellar envelopes of
mass-losing evolved AGB stars
\citep{bressan,clemens}. The strength of this silicate emission is a
slowly declining function of stellar age \citep{piovan}, and persists
even for very evolved stellar populations ($>\!10$\,Gyr). The
optical-MIR colours of Virgo and Coma galaxies have been found to be
consistent with such old stellar populations \citep{clemens}.
\begin{figure}
\centering{{\rotatebox{270}{\epsfig{file=figure-7,width=6.5cm}}}}
\caption{This colour-colour diagram shows that the galaxies identified
as AGN ({\it{green circles}}) and star-forming ({\it{blue stars}}),
on the basis of the BPT diagram according to their optical spectra
(see Fig.~\ref{bpt}), occupy different regions on the near/mid IR
magnitude-colour diagram. The {\it{solid}} and {\it{open green
circles}} represent AGN classified on the BPT diagram
and by the \citet{miller03} criterion respectively (see text). The
passive galaxies ({\it{red points}}) detected at 24{$\mu$m}\ are
concentrated in a small region on the top right of the diagram. The
horizontal line at $(24\!-\!z)= -6$~mag marks the boundary between these
two classes in the $z$-$(24\!-\!z)$ IR colour magnitude
space (Fig.~\ref{ssf}). }
\label{col}
\end{figure}
The blue galaxies in Fig.~\ref{cmr} have a wide spread in both colour
and magnitude. The inhomogeneity of this class of galaxies becomes
even more clear in Fig.~\ref{col}, where we plot the optical
$(g\!-\!r)$ colour against the near/mid-IR $(24\!-\!z)$ colour. Just
as in Fig.~\ref{cmr}, the passive red galaxies cluster in a small
region of the colour-colour space, but the blue galaxies and
(optical) AGNs span a wide range along both the axes. We note the
clear separation of star-forming (blue stars) and passive (red
points) galaxies in $(24-z)$ colour. The horizontal dashed line in
Fig.~\ref{col} indicates our empirically chosen criterion to separate
the two populations about $(24-z)=-6$. Interestingly, although the
AGN and star-forming galaxies (green and blue symbols respectively)
are classified on the basis of their optical spectra on the BPT
diagram \citep[][also see \S\ref{sf}]{bpt81}, they occupy distinct
regions in this plot of optical vs optical-IR colour, as well.
\begin{figure}
\centering{
{\rotatebox{0}{\epsfig{file=figure-8,width=9.5cm}}}}
\caption{The BPT diagram \citep{bpt81} for the 24$\mu$m detected
galaxies in the Coma cluster ({\it{filled circles}}) and in
Abell~1367 ({\it{open triangles}}), colour coded by their 24{$\mu$m}\
flux. We also plot all the galaxies within $\pm$ 2,000 km~s$^{-1}$
of Coma or Abell~1367, in the 500 square degree
supercluster region, but not detected at 24{$\mu$m}, as {\it
grey crosses }. The points plotted in a line at
log[OIII]5007/H$_{\beta}\!=\!-1.5$ are the galaxies which have no
detected [OIII] and/or H$_{\beta}$ emission. It is interesting to
observe that a large fraction of the Coma galaxies without [OIII]
and/or H$_\beta$ have their [NII]/H$_\alpha$ flux ratios as expected
for AGN, suggesting that these galaxies may have their nuclear
emission obscured. We classify such galaxies as AGN if they have
Log([NII]/H$_{\alpha}$)$>$-0.2 \citep{miller03}. }
\label{bpt}
\end{figure}
\subsection{Optical and 24{$\mu$m}\ star formation indicators}
\label{sf}
The understanding of the formation of stars in galaxies requires,
among other things, the measurement of the rate at which the
interstellar gas is converted into stars. With the development of
appropriate technology, radio, IR and UV photometry and spectroscopy
are increasingly being employed, in addition to the traditional
optical observations, for the purpose of measuring the rate of star
formation. The dust clouds surrounding the young stellar
concentrations absorb starlight and re-radiate it at IR
wavelengths. Thus, the SFR measured by IR measurements is accurate
only in the optically thick limit. The observed UV radiation escaped
from the molecular clouds which block the UV light at earlier ages,
comes mostly from stars with ages 10$^7$\,-\,10$^8$ yrs
\citep{calzetti05}. H$\alpha$ emission is produced only in the first
few million years from the most massive stars \citep{leitherer}.
Indeed, a combination of UV, H$_\alpha$ and IR observations is
required to give a full measure of the obscured and unobscured star
formation. In this section, we attempt to combine the direct
(optical) and the obscured (IR) radiation from the Coma and
Abell~1367 galaxies to understand the process of star formation
across this supercluster.
The 24{$\mu$m}\ IR emission in galaxies can result from dust heated by
young massive star clusters as well as the AGN. Since our MIPS data
comes from the densest regions at the core of the Coma cluster
(Fig.~\ref{scl}), which is a favourable environment for AGN hosts, it
is important to investigate the origin of the 24{$\mu$m}\ emission in these
galaxies. To do so, in Fig.~\ref{bpt} we plot the usual ratios of
optical emission linewidths ([OIII], H$_\beta$, [NII] and
H$_\alpha$), known to distinguish star forming galaxies from those
dominated by AGN, in what is popularly known as the BPT diagram
\citep{bpt81}.
In Fig.~\ref{bpt}, we show all the emission-line galaxies found in the
$\pm$\,2,000 km~s$^{-1}$ redshift slice around Coma and/or Abell 1367,
in the $\sim$500 square degrees Coma supercluster region (see
Fig.~\ref{scl}), along with the galaxies detected at 24{$\mu$m}\ (MIPS)
in Coma (circles) and Abell~1367 (open triangles). Although we find
a significant number of 24{$\mu$m}\ detected galaxies with [NII]/H$_\alpha$
ratios suggestive of AGN, this does not mean that the 24{$\mu$m}\ emission
is predominantly due to AGN.
\citet{goulding09} have tried to estimate the contribution of the AGN
component to the IR emission from galaxies using Spitzer/IRS
spectroscopy. Although only a few galaxies in their sample have IR
fluxes produced predominantly by AGN, they also find that a
substantial fraction of AGN are optically obscured, consistent with
the results presented here.
We find that most of the galaxies which have measured values of all
four emission lines are dusty star forming galaxies. However,
interestingly, a large number of galaxies detected at 24{$\mu$m}\ in Coma
that do not show emission in [OIII] and/or H$_\beta$, do have a
[NII]/H$_\alpha$ flux ratio consistent with the presence of an
AGN. Such galaxies can be classified as AGN if
$\log$ ([NII]/H$_{\alpha}$)$>$-0.2 \citep{miller03}.
\begin{figure}
\centering{
{\rotatebox{0}{\epsfig{file=figure-9,width=9cm}}}}
\caption{The ($24\!-\!z$) colour as a function of H$_{\alpha}$ EW for
Coma ({\it{filled circles}}) and Abell 1367 ({\it{stars}}) galaxies
respectively. The straight line represents the linear fit to
galaxies in the Coma cluster. Though there is a considerable
scatter, the correlation between the two quantities is
significant. The dotted lines represent the lower limits adopted to
select star-forming galaxies using (H$_\alpha$ EW=4\AA) and ($24\!-\!z$)
colour\,$=\!-6.0$ (see Fig.~\ref{ssf}). For completeness, we also show the 24{$\mu$m}\
detected galaxies that do not have H$_\alpha$ emission in SDSS spectra at
EW(H$_\alpha$)\,$=\!0.2$\AA~({\it{open circles}}). The 24{$\mu$m}\ emission in these
galaxies is likely to come from the evolved AGB stars. }
\label{ha-24z}
\end{figure}
The EW of the Balmer lines, especially that of the H$_\alpha$
emission line, has been extensively used to estimate the current
optical SFR of galaxies. It has been shown in the literature that the
24{$\mu$m}\ flux is a good measure of dust-processed star light
\citep[{e.g.}][]{calzetti}. But, the 24{$\mu$m}\ observations used in this
work cover the very central regions of the Coma cluster, and
Abell~1367, which comprises mostly of early-type galaxies. In order
to test the correspondence between the direct and dust-processed SFR
tracers, in Fig.~\ref{ha-24z} we plot the $(24\!-\!z)$ colour as a
function of spectroscopically measured EW(H$_\alpha$). We also
colour-code the symbols according to the concentration parameter of
the galaxies (ratio of the Petrosian radii R$_{90r}$/R$_{50r}$ from
the SDSS photometric catalogue, which is an indicator of morphology).
Passive spirals often have concentration indices consistent with early-types
\citep[R$_{90r}$/R$_{50r}\!>\!2.6$; also see][]{mahajan09a}. But
Fig.~\ref{ha-24z} shows that both the spiral and spheroidal galaxies
are uniformly distributed along both the axes, implying that the emission-line
galaxies detected at 24{$\mu$m}\ at the core of Coma and Abell 1367 are not
dominated by galaxies of any particular morphological type (as
quantified by the concentration index).
Due to the dominance of early-type galaxies in our 24{$\mu$m}\ sample, it is
not surprising that only a small fraction of galaxies classified as
star-forming on the basis of their $(24\!-\!z)$ colour
(Fig.~\ref{ssf}), show no signs of current star formation in their
optical spectra.
For completeness, at EW(H$_\alpha$)\,$=\!0.2$\AA~we plot all the 24{$\mu$m}\
detected galaxies that do not show optical emission in H$_\alpha$. As
expected, most of these galaxies have $(24\!-\!z)$ colours and concentration
parameters consistent with those of quiescent early-type galaxies. This
suggests that their 24{$\mu$m}\ emission is primarily due to the contribution
from AGB stars, which (unlike young stars) do not produce H$_\alpha$ emission.
The good overall correspondence between the
photometric $(24\!-\!z)$ colour and spectroscopic EW(H$_\alpha$)
makes the $(24\!-\!z)$ colour a good candidate for comparing the star
formation activity in nearby galaxies, in the absence of optical spectra.
\subsection{Varying fractions of star-forming galaxies with SFR
tracer: implications for the Butcher-Oemler effect}
\label{fractions}
\begin{figure*}
\centering{
{\rotatebox{270}{\epsfig{file=figure-10,width=7cm}}}}
\caption{The fraction of star-forming galaxies as a function of
(projected) cluster-centric radius for Coma in the {\bf{Left}} panel
and Abell 1367 on the {\bf{Right}}. In both panels, the {\it{solid
black line}} shows the radial variation in the fraction of
star-forming galaxies, selected using the $(g\!-\!r)$ colour
threshold, the {\it{dotted red line}} makes use of the H$_\alpha$ EW
($\geq$2\AA) and the {\it{dashed blue line}} represents the same,
for star-forming galaxies chosen by the $(24\!-\!z)\!<\!-6$
criterion. The corresponding {\it{thin curves}} show the $\pm\!1\sigma$
Poisson scatter. In all cases, all the spectroscopically
identified galaxy members with $M_{z}\!<\!-20.82$
($M^*_{z}\!+\!1.5$; see text), and l.o.s. velocity $\pm$ 3,000
km~s$^{-1}$ of Coma and/or Abell 1367, are used to estimate the
fractions. For Abell~1367, we show galaxies only within the region,
of scale $\sim$\,0.75 $h^{-1}_{70}$~Mpc, for which the 24{$\mu$m}\ data is
available. This diagram shows the importance of taking account of
the star formation activity used to quantify and understand
evolutionary trends, such as the Butcher-Oemler effect. For
comparison, the R$_{30}$ radius adopted by \citet{boe}, and the
fraction of blue galaxies found in Coma and Abell 1367, are also
shown by the vertical {\it{dot-dashed lines}} and {\it{open green
diamonds}} respectively.}
\label{sf-frac}
\end{figure*}
Photometric colours and EWs of emission lines like [OII] and
H$_\alpha$ have been extensively used for studying the evolution of
galaxies in time and across the sky. In one such pioneering work,
\citet[][BO84]{boe} found that clusters at moderate to high redshifts contain
an `excess' of blue galaxies, relative to their local counterparts.
They estimated the blue fraction by considering galaxies found within
R$_{30}$ (radius containing 30\% of all red sequence galaxies), for
which the optical broadband colours are bluer by at least 0.2 mag
than that of the red sequence galaxies. For the redshift regime of
Coma ($\leq\!0.1$), BO84 found a uniform blue fraction within
R$_{30}$. Since then, several such studies
have sought to quantify
and validate the Butcher-Oemler effect for different samples
of clusters. Some of these
define their samples in a way similar to that of BO84,
and obtain similar results
\citep[{e.g.}][]{margoniner00}.
Using panoramic MIR data for 30
clusters (including both Coma and Abell 1367) over $0\!<\!\rm z\!<\!0.4$,
\citet{chris09} are able to reproduce the Butcher-Oemler effect using
a fixed limit in $L_{IR}$ $(5{\times}10^{10}L_{\odot})$ (equivalent
to a fixed SFR of 8\,M$_{\odot}$yr$^{-1}$), but show that the
Butcher-Oemler effect can be largely explained as a consequence of
the {\em cosmic} decline in star formation \citep{lefloch,zheng07}.
In this case, the blue galaxies in clusters are those that are recently
accreted from the field \citep[accretion has occurred at a relatively constant
rate since $\rm z\!\sim\!0.5$;][]{berrier09}, but since the global
level in star formation among these galaxies has declined, a smaller
fraction of the infalling population is classed as blue (which
assumes a non-evolving level of star formation), resulting in the
observed Butcher-Oemler effect.
Elsewhere \citep{bnm00,ellingson01,depropris04}, studies going out to
several multiples of the cluster-centric radius, scaled by R$_{200}$,
show the effect of the chosen aperture size on the blue fraction. It
has been shown that the observed gradual radial trend of $f_{SF}$ is
consistent with a simple infall scenario, whereby the star-forming
galaxies are infalling field galaxies, which are then quenched
rapidly upon their first passage through the cluster core
\citep[{e.g.}][]{balogh00,ellingson01,chris09}. By comparing the
photometric colour with the spectroscopically determined SSFR,
\citet{mahajan09a} show that the presence of metal-rich stellar
populations in low redshift cluster galaxies can also influence the
evolutionary trends such as the Butcher-Oemler effect, if they are
studied only using galaxy colours. Several recent studies indicate
that a non-negligible fraction of red sequence galaxies show signs of
ongoing star formation from their optical spectra and/or broad-band
colour, and that a robust separation of passive and star-forming
galaxies requires mutually independent data
\citep[{e.g.}][]{bildfell}. In this work, we show that in the
Coma Supercluster,
the value of $f_{SF}$ not only varies with the
cluster-centric aperture used for measuring the blue fraction, but
can also be severely effected by the SFR tracer employed.
In Fig.~\ref{sf-frac} we plot the fraction of non-AGN, star-forming
galaxies found using 3 different criteria: (i) IR colour
[($24\!-\!z)\!\leq\!-6$ mag; Fig.~\ref{ssf}], (ii)
EW(H$_\alpha$)$\geq\!2${\AA}~and, (iii) the photometric colour
$(g\!-\!r)$ is bluer than that of the fitted red sequence
(Fig.~\ref{cmr}, by more than the mean absolute deviation, for all
the Coma and Abell 1367 galaxies. BO84 had included galaxies brighter
than $M_{V}\!=\!-20$ (H$_{0}\!=\!50$ km~s$^{-1}$Mpc$^{-1}$) in their
sample. To make a fair comparison with the work of \citet{boe},
without adding uncertainties by using empirical relations to convert
magnitudes from one passband to the other, in Fig.~\ref{sf-frac} we
choose to only include galaxies brighter than $M^{*}_{z}\!+\!1.5$
\citep[$M^{*}_{z}\!=\!-22.32$;][]{blanton01} for calculating the
fractions. Also, note that {\it all} the spectroscopic galaxy members
are used to calculate the fractions in each radial bin.
These distributions show that, by taking into account the obscured
star formation estimated from the 24{$\mu$m}\ flux, the fraction of
star-forming galaxies can dramatically vary at any given radius from
the centre of the cluster. In both clusters, the `blue' fraction
($f_b$) is higher, and flattens at lower cluster-centric radii, than the
$f_{SF}$ estimated using the
EW(H$_\alpha$), evidently showing a significant contribution of the
post-starburst galaxies to the fraction $f_b$ (also see
Section~\ref{ka}). We note that this trend may vary if the dusty red
galaxies have a non-negligible contribution
in building the red sequence. But as seen in Fig.~\ref{cmr}, in the
Coma cluster, this does not seem to be the case.
It is interesting to note that unlike Coma, Abell~1367 has an
increasing fraction of blue galaxies with cluster-centric radius
(black lines), but an inverse trend emerges when the fractions are
measured using the $(24\!-\!z)$ colour (blue lines) or the H$_\alpha$
EW (red lines). This implies that some of the blue galaxies outside
the core of Abell~1367 are post-starburst galaxies (see
Fig.~\ref{coma-gals}). This result supports the results obtained in
the more general work of \citet{mahajan09a}, who show that using a
single galaxy property, such as the broadband colour, is not an
appropriate way of quantifying evolutionary trends like the
Butcher-Oemler effect.
\section{Discussion}
\label{discussion}
In this paper we set out to understand the relationship between the
star formation activity in galaxies, as depicted by an assortment of
indicators encompassing the optical and 24{$\mu$m}\ mid-IR wavebands, and
their immediate and global environment, in the Coma supercluster. A
wide range of local and global environments of galaxies, and a
uniform optical coverage across the entire $\sim\!500$ square degrees
of sky, make this supercluster an ideal laboratory for examining the
environmental dependence of galaxy properties. We discuss below the
implications of the results from our analysis presented in
\S\ref{coma-sf} and \S\ref{results}.
\subsection{The spatial and velocity distribution of galaxies detected
at 24{$\mu$m}}
\label{mips}
\begin{figure}
\centering{
{\rotatebox{270}{\epsfig{file=figure-11,width=9.0cm}}}}
\caption{{\bf{(Top panel):}} The
distribution of cluster-centric distance for all the
spectroscopically identified
galaxies in a $\pm$3,000 km~s$^{-1}$ slice in the Coma cluster
({\it{solid black}} histogram).
The {\it{dashed red}} and {\it{dotted blue}} histograms show the
same for the 24{$\mu$m}\ MIPS galaxies, divided into
red and blue galaxies on the basis of the
$(g\!-\!r)\!-\!r$ colour-magnitude plot (Figs.~\ref{cmr} \&
\ref{ssf}). The horizontal axis is limited by the coverage of the
MIPS field. Interestingly, even though the red 24{$\mu$m}\ galaxies are
uniformly distributed within a $\sim$2~Mpc radius, the blue galaxy
population seems to peak away from the cluster core at 1--1.5~Mpc from
the centre. {\bf{(Bottom panel):}} These histograms represent
the same galaxies as in the upper panel, but for the
l.o.s. velocity of galaxies, relative to the mean redshift of the
Coma cluster. Intriguingly, while the red galaxies follow the
distribution of {\it all} the spectroscopic galaxy members of the Coma
cluster, the blue 24{$\mu$m}\ galaxies show a remarkable peak around the
velocity of the galaxy group NGC\,4839, shown here as the
{\it{dot-dashed}} line, with $\pm \sigma_v\!=\!329$\,km~s$^{-1}$
\citep{colless96}. }
\label{radius}
\end{figure}
The 24{$\mu$m}\ data for the Coma cluster extend out to a few times its
core radius (Fig.~\ref{sf-frac}), allowing us to analyse the spatial
(sky and velocity) distribution of the galaxies detected at 24{$\mu$m},
relative to all the spectroscopic members found in the SDSS. In order
to do so, in Fig.~\ref{radius}, we plot the distribution of all the
spectroscopic galaxies, and the (optically) red and blue galaxies
detected at 24{$\mu$m}\ respectively. As discussed above (Figs.~\ref{cmr}
and \ref{ssf}), in Coma both the 24{$\mu$m}\ and optical colours mostly
segregate the same galaxies into `red' and `blue'
ones. Fig.~\ref{radius} (top panel) shows that, of the 24{$\mu$m}\ detected
galaxies, the distribution of the (optically) red ones is similar to
that of all the galaxies, but the (optically) blue ones tend to peak
$\gtrsim$1.0 $h^{-1}_{70}$Mpc from the centre of Coma.
In Fig.~\ref{radius} (bottom panel), we plot the distribution of the
line of sight (l.o.s.) velocities of galaxies, with respect to the
mean velocity of the Coma cluster (called the `relative velocity'
hereafter), for the same three sets of galaxies. The distribution of
the relative velocity of `all' and (optically) red galaxies are
statistically similar. But the (optically) blue 24{$\mu$m}\ galaxies show a
bimodal distribution, with a large fraction of one mode concentrated
around relative velocity $\sim\,600$ km~s$^{-1}$. A Kolmogorov-Smirnoff
(K-S) test suggests
that the probability of the parent distribution of the relative
velocities of red and blue galaxies being sampled from the same
population is $p\!=\!8.023\times10^{-5}$. The same comparison
between the relative velocity distribution of `all' and blue galaxies
gives $p\!=\!0.005$, suggesting that it is highly unlikely that the
blue and red (or `all') galaxies are drawn from the same parent
distribution (Fig.~\ref{radius}). This leads us to conclude that a
non-negligible fraction of the (optically) blue 24{$\mu$m}\ galaxies within
$2\,h^{-1}_{70}$~Mpc of the centre of the Coma cluster are
star-forming galaxies which may belong to the substructure associated
with NGC\,4839 (see \S\ref{ngc4839} for further discussion).
The l.o.s. velocity distribution of the (optically) blue 24{$\mu$m}\ galaxies
is highly non-Gaussian, with an excess at $\sim\!600$ km s$^{-1}$, suggesting
that many of them have recently entered the cluster. If these blue 24{$\mu$m}\
galaxies were virialized, a Gaussian distribution centered on $0$,
like that seen for the red (red dashed histogram) and `all' (solid black histogram)
galaxies would be expected.
The current episode of star formation in these galaxies can be
attributed to the environmental impact of the cluster's ICM
\citep[{e.g.}][]{poggianti04}, or the enhanced galaxy density in the
infall region \citep[{e.g.}][]{mrp10}. This result is consistent with
the findings of \citet{caldwell97}, who analysed the spectra of
early-type galaxies in 5 nearby clusters, including Coma, and found
that in 4 of the 5 clusters, the early-type galaxies show
signatures of recent star formation in their spectra (also see \S\ref{ka}).
\subsection{The distribution of k+A galaxies}
\label{ka}
\begin{figure*}
\centering{
{\rotatebox{270}{\epsfig{file=figure-12ap,width=6.5cm}}}}
\centering{
{\rotatebox{270}{\epsfig{file=figure-12b,width=6.5cm}}}}
\caption{{\bf{(Left panel):}} The distribution of the passive {\it{(red
dots)}}, AGN host {\it{(green points)}}, star-forming {\it{(blue
stars)}} and the post-starburst (k+A) {\it{(purple circles)}}
galaxies in $\sim\!5.0\!\times\!4.2$\,$h^{-1}_{70}$Mpc region
surrounding the centre of Coma. The contours show X-ray emission
from a {\it{XMM-Newton}} EPIC/PN observation.
As can be noticed, not only the
star-forming, but also the k+A galaxies seem to avoid the dense
cluster centre. Also, the presence of k+A galaxies out to almost
twice the virial radius from the centre shows how strong the impact
of the `global' cluster environment is on the evolution of galaxies
in the vicinity of massive structures.
{\bf{(Right panel):}} Same as above, but for the
$5.0\!\times\!5.0$\,$h^{-1}_{70}$Mpc region surrounding the centre
of Abell 1367. It is interesting to see that in contrast with Coma,
most of the k+A galaxies in Abell 1367 seem to be aligned along the
direction of the filament feeding it from the direction of Coma. We
also note that the (optical) AGN in both the clusters are mostly
found in the direction of the filament connecting Coma and Abell
1367. }
\label{coma-gals}
\end{figure*}
\begin{figure}
\centering{
{\rotatebox{270}{\epsfig{file=figure-13,width=6.5cm}}}}
\caption{ The $(g\!-\!r)\!-\!z$ colour-magnitude diagram of all the
spectroscopically identified galaxies {\it{(black points)}} in the
$\sim\!5.0\!\times\!4.2$\,$h^{-1}_{70}$Mpc region surrounding the
centre of the Coma cluster (shown in Fig.~\ref{coma-gals}).
The k+A galaxies {\it{(purple diamonds)}} are mostly blue
dwarfs. These observations confirm that the low-mass galaxies are
the first ones to experience and reflect the impact of rapid change in
their local environment. }
\label{ka-cmr}
\end{figure}
The post-starburst, or k+A galaxies, as they are popularly known in
the literature, are passive galaxies that have recently (1-1.5 Gyr)
experienced a strong burst of star formation. These galaxies are very
crucial in understanding the impact of environment on various galaxy
properties, especially their SFR. The spectrum of a k+A galaxy shows
strong absorption in H$_\delta$, but little or no emission in
H$_\alpha$.
In this work we make use of the SDSS DR7 spectroscopic galaxy
catalogue ($r\!\leq\!17.77$) for identifying the k+A galaxies
(EW(H$_{\delta})\!<\!-3$\AA~\& EW(H$_{\alpha})\!<\!2$\AA) in the Coma
supercluster (Please note that throughout this work, negative values
of EW indicate absorption). In Fig.~\ref{coma-gals} we show these k+A
galaxies together with the passive, star-forming and AGN galaxies in
and around the Coma and Abell 1367 cluster respectively. As can be
easily seen, the k+A galaxies preferentially avoid the dense region
in the cluster core but inhabit the surrounding infall regions out to
$\sim\!5$\,$h^{-1}_{70}$Mpc (almost twice the virial radius). In
Fig.~\ref{coma-gals} we have also over-plotted the contours
of intensity from the mosaicked 0.5--2~keV image of the core of Coma
cluster from XMM-Newton EPIC/PN observations \citep{fino03}, kindly
supplied to us by A.~Finoguenov. We note that there are almost no
star-forming galaxies in the X-ray emitting region of the core. This
observation further strengthens the argument that the changes in
environment, experienced by a galaxy on the outskirts of clusters,
play a key role in modulating the properties of galaxies, especially
the dwarfs \citep[{e.g.}][]{porter07,porter08}.
By analysing deep ($M_{B}\!\lesssim\!-14$) photometric and
spectroscopic optical data for 3 regions in Coma (two near the centre
and one near NGC\,4839, each $\sim\!1\!\times\!1.5$ Mpc in size),
\citet{poggianti04} found that $\sim$\,10\% of the cluster's dwarf
($M_{V}\!>\!-18.5$) galaxies have post-starburst spectra. They also
used the results from the X-ray analysis of \citet{neumann03} to show
that the k+A galaxies in the Coma cluster are likely to be a result
of the interaction between the Coma cluster and the adjoining
NGC\,4839 galaxy group (see \S\ref{ngc4839}). In
Fig.~\ref{coma-gals}, we overplot as dashed rectangles, the
approximate location of the substructures found by
\citet{neumann03}. As has been shown by \citet[][their
fig.~6]{poggianti04}, on the western side of the Coma cluster, the
k+A galaxies seem to lie along the X-ray substructure \citep[also see
fig.~2 of][]{neumann03}. We demonstrate this by overplotting contours
of X-ray intensity in Fig.~\ref{coma-gals}.
However, it is also interesting to note that the southern substructure,
apparent from the X-ray contours, is almost devoid of k+A galaxies,
but has a stream of star-forming dwarf galaxies flowing towards the
cluster core (Fig.~\ref{mean_ha}; also see \S\ref{coma-scl}). Contrary to
\citet{poggianti04}, by taking into account the extended region surrounding
the Coma cluster, we find that the k+A galaxies preferentially avoid the
cluster core but their spatial distribution does not show any correlation
with the substructure evident in the X-ray emission (Fig.~\ref{coma-gals}).
In Fig.~\ref{ka-cmr} we examine the position of the k+A galaxies in
the Coma cluster in the plot of $(g\!-\!r)$ colour vs $z$ magnitude.
Fig.~\ref{ka-cmr} show that almost all the k+A galaxies in Coma are
dwarfs ($z\!\lesssim\!15$), suggesting that either only the infalling
dwarf galaxies are rapidly quenched, or the SFR-$M^*$ relation
\citep[{e.g.}][]{feulner} dilutes the post-starburst signature in
massive galaxies.
Since dwarf k+A galaxies in the Coma supercluster are found on the
outskirts of Coma and Abell~1367 clusters, and occasionally in galaxy
groups embedded elsewhere in the large-scale structure (LSS), this
might suggest that dwarf galaxies falling into deeper potentials are
more likely to show k+A features, while those being assimilated into
galaxy groups may be quenched on a longer time-scale. We will probe
this in greater detail in a later paper.
The handful of galaxies with very blue colours ($(g\!-\!r)\!<\!0.5$;
Fig.~\ref{ka-cmr}) are the k+A dwarfs in which the episode of
starburst could have ended $\sim\!300$\,Myr ago, unlike the more
evolved red post-starburst dwarfs \citep{poggianti04}. By studying
spectra of early-type galaxies ($M_{b}\!<\!16.7$) in Coma,
\citet{caldwell97} also found that $\sim$\,15\% of these galaxies
show signatures of recent or ongoing star formation. Based on this
result, \citet{caldwell97} suggested that the present day clusters
act as a catalyst for galaxy evolution, though at a reduced level as
compared to their high redshift counterparts.
\begin{table}
\begin{minipage}{\linewidth}
\centering{
\caption{Dwarf ($z\!>\!15$ mag) galaxies in the Coma supercluster}
\label{tbl:ka-nos}
\begin{tabular}{|c|c|c|c|}
\hline
& Total & Star-forming & k+A \\
\hline
Coma & 438 & 55 & 50 \\
($\sim\!5\!\times\!4.2$\,Mpc$^2$; Fig.~\ref{coma-gals}) & & & \\
Abell 1367 & 391 & 203 & 19 \\
($\sim\!5\!\times\!5$\,Mpc$^2$; Fig.~\ref{coma-gals}) & & & \\
Supercluster & 1106 & 667 & 23 \\
(excluding regions mentioned above) & & & \\
\hline
\end{tabular}}
\end{minipage}
\end{table}
Given the vulnerable nature of dwarf galaxies, it is not surprising
that a large fraction of them are the first ones to reflect the
impact of rapid changes in their environment during the transition
from filament to cluster. This contributes to dwarf galaxies having a
better-defined SF-density relation, relative to their more massive
counterparts (Fig.~\ref{mean_ha}).
Table~\ref{tbl:ka-nos} shows that
11.4\% of dwarfs in Coma, 4.8\% in Abell 1367 and 2.1\% in the
neighbouring supercluster region have spectra with k+A features. This
evidently shows that the mechanisms responsible for quenching star
formation in dwarf galaxies and rapidly transforming them to passive
galaxies via the post-starburst phase, are strongly dependent on the
cluster potential. This result is also in agreement with the
observation of excessive red, dwarf ellipticals (dEs) in the Coma
cluster \citep{jenkins07}. In Table~\ref{tbl:ka-catalogue} we provide a
list of all the 110 k+A dwarf ($z\!>\!15$) galaxies found in the
entire Coma supercluster.
\begin{table}
\begin{minipage}{\linewidth}
\centering{
\caption{Catalogue of dwarf ($z\!>\!15$),
k+A (EW(H$_\alpha$)\,$<\!2$\,\AA~ \&
EW(H$_\delta$)\,$<\!-3$\,\AA) galaxies in the Coma supercluster (this table is
available in full online).}
\label{tbl:ka-catalogue}
\begin{tabular}{|c|c|c|c|c|c|}
\hline
RA & Dec & Redshift & $z$ & EW(H$_\alpha$) & EW(H$_\delta$) \\
(J2000) & (J2000) & & mag & \AA & \AA \\
\hline
195.4972 & 28.7095 & 0.0203 & 16.8559 & -1.9535 & -20.5023 \\
194.7338 & 28.4636 & 0.0198 & 15.1591 & -2.6499 & -5.6373 \\
174.6209 & 28.5871 & 0.0236 & 15.3615 & 1.8792 & -3.9275 \\
192.4220 & 28.8448 & 0.0218 & 17.1838 & -1.3320 & -30.0913 \\
192.8569 & 28.7203 & 0.0241 & 17.0417 & -1.1904 & -3.1975 \\
193.9028 & 26.5644 & 0.0211 & 16.9923 & -1.8170 & -6.2012 \\
194.5201 & 26.3706 & 0.0232 & 15.7774 & -1.8387 & -6.3831 \\
195.0435 & 26.4612 & 0.0221 & 16.6272 & -2.0973 & -3.2747 \\
189.0882 & 27.0333 & 0.0251 & 16.6488 & -1.9006 & -3.0789 \\
194.0083 & 26.9208 & 0.0191 & 16.5278 & -1.4534 & -4.9134 \\
\hline
\end{tabular}}
\end{minipage}
\end{table}
\subsection{Coma and NGC\,4839}
\label{ngc4839}
\begin{figure}
\centering{
{\rotatebox{270}{\epsfig{file=figure-14p,width=6.5cm}}}}
\caption{The spatial distribution of galaxies detected at 24{$\mu$m}\ in the
Coma Supercluster region. The galaxies are coded by their
($24\!-\!z$) colour as following: red:$>\!-4$, orange:
$-4\leq\!(24\!-\!z)\!<\!-6$, green:$-6\leq\!(24\!-\!z)\!<\!-8$ and
blue: $\leq\!-8$. The {\it{purple circles}} are the blue 24{$\mu$m}\ galaxies
moving at velocities similar to that of the NGC\,4839 galaxy group
\citep[{i.e.}~having l.o.s.
velocity\,$=\!7339\!\pm\!329$\,km~s$^{-1}$;][]{colless96}. We note
that 19 of these 26 galaxies have ($24\!-\!z$) colour corresponding
to that of the passive galaxies ({i.e.}~$(24\!-\!z)\!\leq\!-6$;
Fig.~\ref{ssf}). The contours are from a 0.5--2~keV XMM-Newton
EPIC/PN X-ray mosaic image of the extended Coma cluster. }
\label{coma-mips}
\end{figure}
As noted above, the Coma cluster is known to have significant
substructure in the optical and X-ray maps. The most prominent X-ray
emitting substructure is associated with a galaxy group, of which
NGC\,4839 is the most prominent galaxy. By combining N-body
hydrodynamical simulations with {\it{ROSAT}} X-ray and VLA radio
observations, \citet{burns94} remarked that the NGC~4839 group has
already passed through the core of Coma $\sim$\,2 Gyr ago, and is now
on its second infall \citep[also see][]{caldwell97}. However, several
other studies based on optical spectroscopy and imaging, and X-ray
observations conclude otherwise
\citep{colless96,neumann03,adami05}. No consensus seems to have
been reached on the dynamical state of NGC\,4839. Studies supporting
the first infall argument suggest that the NGC\,4839 galaxy group has
a velocity dispersion ($\sigma_{cz}$) of 329 km~s$^{-1}$
\citep[{e.g.}][]{colless96}, while others claim it to be as high as 963
km~s$^{-1}$ \citep[{e.g.}][]{caldwell93}, the latter being consistent
with the scenario where the galaxies belonging to the group have
dispersed during their first passage through the cluster.
Located $\sim\,1.1$~Mpc from the centre of the Coma
cluster, the NGC\,4839 galaxy group was first detected as an
asymmetric extension to the otherwise relaxed X-ray morphology of
Coma \citep{briel92}. Several authors have discovered post-starburst
(k+A) galaxies associated with this group
\citep{caldwell93,caldwell97,poggianti04}. In this work, we find
that most of the known post-starburst galaxies in Coma, such as those
found by \citet{poggianti04}, are either not detected at 24{$\mu$m}, or show
little MIR emission. We hence confirm that almost none of these
post-starburst galaxies have a significant amount of obscured star
formation going on in them.
\citet{struck06} suggests that when a galaxy group passes through the
core of a cluster, as suggested by \citet{burns94} for Coma and
NGC\,4839, the group galaxies are gravitationally shocked. In this
scenario, enhancement in star formation activity is a natural
consequence of an increase in galaxy-galaxy interactions among group
galaxies. Such a scenario has been suggested for Coma and NGC\,4839
by \citet{neumann03}. Fig.~\ref{coma-mips} may provide
circumstantial evidence to support this argument.
In Fig.~\ref{coma-mips} we show the spatial distribution of the 24{$\mu$m}\
detected Coma galaxies colour-coded by their ($24\!-\!z$) colour. The
blue 24{$\mu$m}\ galaxies constituting the peak, which also corresponds to
the mean velocity of the NGC\,4839 group (Fig.~\ref{radius}), are
explicitly shown.
However, poor statistics in our data do not allow us to favour any of
the 2 scenarios associated with the NGC\,4839 galaxy group and the Coma
cluster, namely, (i) whether the group is on it's first infall, or (ii)
it has already passed through the core and is on it's second passage.
\subsection{Coma, Abell 1367 and the filament}
\label{coma-scl}
In the hierarchical model, galaxy clusters grow by accretion and/or
mergers with other clusters and groups. In simulations,
galaxies are seen to be accreted along
the network of filaments of galaxies, feeding these clusters,
along preferred directions
\citep[{e.g.}][]{bkp96}. Although several
observational techniques are now being implemented to detect and
quantify such large-scale structures (LSS), the low surface
density of matter, enormous spatial scale, and projection
effects make it difficult to observe the extent of
these filaments of galaxies \citep{colberg07}.
One way of exploring the evolution of the LSS is to probe its impact
on the galaxies traversing them. The spectacular filament crossing
the Coma and Abell~1367 clusters is an exclusive object in the low
redshift Universe, because it is not only traced by the spatial
distribution of galaxies (Fig.~\ref{scl}), but has also been detected
at radio wavelengths \citep{kim89}. One of the aims of this paper is
to understand and interpret the difference in the evolutionary paths
adopted by galaxies in the vicinity of the clusters, the filament and
the groups embedded in the filament. In agreement with the scenario
emerging in the literature \citep[][among
others]{gray04,tanaka,smith06,chris07}, we find that the SFR-density
relation across the Coma supercluster is much weaker for the massive
galaxies ($z\!<\!14.5$; Fig.~\ref{sf_frac}), relative to the dwarfs
($z\!>\!15$; Figs.~\ref{sf_frac} \&~\ref{mean_ha}).
Empirical studies of galaxies infalling into clusters along filaments
suggest that an enhanced galaxy density at the cluster periphery may
lead to a burst of star formation in them, consuming a large fraction
of cold gas \citep[{e.g.}][]{porter08,mrp10}. This effect would be more
efficient for the dwarf galaxies. In a study based on the SDSS data,
\citet{chris07} find that the dwarf galaxies ($-19\!<\!M_r\!<\!-18$)
residing in high density regions show a systematic reduction of
$\sim\!30\%$ in their H$_\alpha$ emission relative to the mean of the
sample, leading the authors to favour slow quenching of star
formation in these galaxies \citep[also see][]{balogh04a,tanaka}. On
the contrary, in Coma we find that the infalling star-forming dwarf
galaxies undergo a burst of star formation, followed by rapid
quenching (Fig.~\ref{mean_ha}, also see Section~\ref{ka}).
In contrast, star formation in dwarf galaxies infalling into galaxy
groups, seems to be slowly quenched without an intermediate starburst
phase, resulting in the observed reduction in the mean EW(H$_\alpha$)
values seen in the vicinity of groups in Fig.~\ref{mean_ha}. This
observation could be attributed to the relatively inefficient
ram-pressure stripping in galaxy groups \citep{tanaka}, transforming
infalling galaxies slowly over several Gyr via a process involving
progressive starvation \citep{larson}.
Another notable fact in Fig.~\ref{mean_ha} is the orientation of the
stream of blue star-forming galaxies southward of Coma, almost
orthogonal to the direction of elongation of the galaxy density
(Figs.~\ref{sf_frac} \& \ref{mean_ha}). This may be significant given
the observations presented in the previous section \citep[also
see][]{burns94}. Consistent with the results presented in
literature, we find that the instantaneous SFR of a galaxy
depends upon the stellar mass of the galaxy, as well as on the
local galaxy density, and on whether the galaxy
is in a group or cluster.
The cosmic web does play a crucial role in defining the
evolutionary path of the galaxy, as is seen by different rate of
quenching of dwarf galaxies in groups and clusters
(Fig.~\ref{mean_ha}). The different fractions of dwarf k+A galaxies
in the three major components of this supercluster further strengthens
this argument (Table~\ref{tbl:ka-nos}). It is indeed remarkable that
the Coma and Abell~1367 clusters, that are well known to have very
different galaxy properties, influence the star formation
histories of the infalling galaxies very
differently. While the fraction of star-forming and k+A dwarf galaxies
in Abell 1367 vary by a factor of 10 ($\sim$\,52\% \& 4.8\%
respectively; Table~\ref{tbl:ka-nos}), in the Coma cluster neither
dominate (12.5\% \& 11.4\% respectively; Table~\ref{tbl:ka-nos}).
\section{Conclusions}
\label{conclusions}
This work is a step ahead in understanding the star formation
properties of galaxies in one of the richest nearby superclusters. We
analyse the spectroscopic and photometric data obtained by the SDSS
and archival 24{$\mu$m}\ data obtained by the MIPS instrument on board {\em
Spitzer}. Our major results are:
\begin{itemize}
\item The fraction of (optical) AGN drops significantly
($f_{AGN}\!<\!0.25$) in the dense cluster environment. But the
relation between AGN activity and environment is unclear in the
intermediate density environments of galaxy groups.
\item Star formation in massive galaxies ($z\!<\!14.5$) seems to
be low everywhere in the supercluster region studied here, almost
independent of the local environment. In sharp contrast, the
dwarf galaxies ($z\!>\!15$) can be seen to be rapidly forming stars
everywhere, except in the dense cluster and group environments.
\item The passive, AGN host and star-forming galaxies as classified
from their optical spectra, occupy different regions on the
$(24\!-\!z)\!-\!(g\!-\!r)$ colour-colour diagram.
\item The fraction of star-forming galaxies in Coma is determined to
within $\pm 10$\%, irrespectively of the definition of star-forming
in terms of optical $(g\!-\!r)$ colour, optical-mid-IR $(24\!-\!z)$
colour or EW(H$_\alpha$). Many of the blue galaxies in Coma are
found to be post-starburst galaxies, whose blue colours are due to
a recent burst of star formation which has now terminated, as
revealed by their lack of H$_\alpha$ emission and excess H$_\delta$
absorption. However, in Abell~1367, the $f_{SF}$ obtained using
the 3 different indicators show different trends. While the
fraction of blue galaxies increases outward from the centre, the
$f_{SF}$ obtained by employing the ($24\!-\!z$) near/mid IR colour
decreases away from the centre of Abell 1367.
\item Most of the (optically) blue 24{$\mu$m}\ galaxies detected in Coma are on their
first infall towards the cluster. The current episode of star
formation in such galaxies is possibly a result of a rapidly changing local
environment.
\item 11.4\% of all dwarf ($z\!>\!15$) galaxies within
$5\!\times\!4.2$~$h^{-1}_{70}$~Mpc$^2$ of the centre of Coma, and
4.8\% within the same area around Abell 1367 have post-starburst
(k+A) type spectra. In the surrounding supercluster region this
fraction drops to 2.1\% only, suggesting that the mechanism(s)
responsible for quenching star formation in dwarfs depends upon the
cluster's potential. The starburst, rapid quenching and subsequent
k+A phase requires the dense ICM and high infall velocities attainable in
rich clusters, as opposed to galaxy groups where star formation in
infalling dwarf galaxies appears to be quenched gradually. The k+A
galaxies preferentially avoid the dense centre of the cluster.
\item The spatial distribution of the k+A galaxies suggests a
correlation between the substructure of the Coma cluster (revealed
in the X-ray emission) and the mechanisms responsible for quenching
star formation in galaxies.
\end{itemize}
\section{Acknowledgments}
We thank Dr Alexis Finoguenov and Dr Ulrich Briel for providing us the
XMM-Newton EPIC/PN 0.5--2~keV mosaic image of the Coma cluster used
in Figs.~\ref{coma-gals} and \ref{coma-mips}. SM is supported by
grants from ORSAS, UK, and the University of Birmingham. CPH
acknowledges financial support from STFC.
We are very grateful to the anonymous referee for constructive comments
that were very helpful in improving this paper.
This research has made use of the SAO/NASA Astrophysics Data System,
and the NASA/IPAC Extragalactic Database (NED).
Funding for the SDSS and SDSS-II has been provided by the Alfred
P. Sloan Foundation, the Participating Institutions, the National Science
Foundation, the U.S. Department of Energy, the National Aeronautics and Space
Administration, the Japanese Monbukagakusho, the Max Planck Society, and the Higher
Education Funding Council for England. A list of participating institutions
can be obtained from the SDSS Web Site http://www.sdss.org/.
|
\section{Introduction}
Though the theoretical study of two-dimensional carbon has a long
history \cite{wall},\cite{Slon},\cite{divinch},\cite{ando1} only
after experimental evidence of existence of graphene as a stable
two-dimensional crystal \cite{alpha},
\cite{berg,geim,zhang} this material became very
popular. The presence of zero gap and zero electron mass,
combined with a rather high mobility at room temperature,
makes graphene an unique material for various fundamental and applied
problems. At present graphene is intensively studied both theoretically
and experimentally (see e.g. reviews
\cite{geim2,rev4}).
The study of graphene optics (see \cite{alpha3},\cite{falk}) is
stimulated by the prediction that the absorption in monolayer
graphene should be determined by the fundamental constant
$\alpha=e^2/\hbar c$ \cite{ando}, \cite{alpha2} and its
experimental evidence \cite{nair}. The investigation of coupling
between photons and electrons in graphene attracts now an active
interest of the community (see e.g. \cite{ziegler1,ziegler2}). An
observation of amplified stimulated terahertz emission from
optically pumped epitaxial graphene heterostructures has been
reported recently \cite{thzgener}. However, the photoinduced
currents, namely, photon drag and photogalvanic effects in
graphene were beyond of interest of the researchers. In this paper
we present the theoretical analysis of these effects.
The study of light pressure on solids has rather long history.
The simplest variant is an instantaneous transmission of photon
momentum to electrons. This process is permitted for interband
transitions or in presence of the ''third body'', for example,
phonons, other electrons, impurities. For a free particle this
process is forbidden by conservation laws. Small value of the
photon momentum makes Nonresonant photon Drag Effect (NDE) extremely weak.
At the same time there exists a less known variant of this effect,
namely Resonant photon Drag Effect (RDE) which has no weakness of
usual NDE
\cite{grinb},\cite{kast},\cite{alper}. Resonance drag occurs when
some partial kinetic property of electron gas sharply depends on
electron energy. A small photon momentum gives an
increase of the electron energy, that can drastically change the
relaxation time. This leads to a significantly
different contributions to the electron current for
electrons exited along or oppositely to the photon direction.
In \cite{alper} the situation was studied for interband transitions
in weakly doped GaAs when the electron energy approaches the
energy of longitudinal optical phonon. In this case electrons
exited along the direction of photon have larger energy than
electrons in opposite direction. Hence, their energy can exceed
the threshold for emission of optical phonon: they quickly emit
phonons and stop, while the opposite electrons will move freely
till they collide with impurity. This gives rise to the appearance of
charge flow in the direction opposite to the light ray.
Here we develop another idea for RDE based on a sharp Fermi
distribution which forbids the transitions below the Fermi energy
$\epsilon_F$. This idea is illustrated in Fig.1. Electrons are
excited from the hole cone to the electron cone by photons with
frequency $\omega$ and wave vector ${\bf Q}$. The conditions for
resonant transitions are $sk+s|{\bf k-q}|=\omega$, $\hbar s|{\bf
k}|>\epsilon_F$, where ${\bf k}$ is the electron momentum counted
from the cone point, $s\approx 10^8$cm/s is the electron velocity,
and ${\bf q}$ is a projection of the wave vector ${\bf Q}$ of
radiation to the plane of graphene. The first condition determines
ellipse in ${\bf k}$ plane, the second limits a part of this
ellipse accessible for transitions. The wave vector tilts the
transitions towards its direction. Fig.~1 shows the case when the
frequency is close to $2\epsilon_F$. The electrons in the figure
are excited from the right segment of the Fermi surface contour.
This results in electron flow rightwards. Since $q\ll k_F$ the RDE
appears when the frequency is close to $2\epsilon_F$, namely if
$|\omega-2\epsilon_F|<sq$. Inside this window the current of RDE
has no smallness connected with $q$ and can be estimated as $j\sim
es\tau ~\pi\alpha~ P/(\hbar\omega)$, where $e$ and $s$ are the
electron charge and the velocity, $\pi\alpha$ is the opacity of
graphene, $\tau$ is the transport relaxation time and $P$ is the
light intensity. Physical meaning of this estimation is evident:
$\tau \pi\alpha P/(\hbar\omega)$ is the instantaneous density of
exited electrons which conserve their momentum. Being multiplied
by the current of individual electrons $es$, this quantity gives
the current density.
Below we determine both NDE and RDE
for interband transitions in monolayer graphene with degenerate
electron gas. Due to graphene electron-hole symmetry results are
applicable to n- and p-type graphene. In general the relaxation process for
electrons and holes are different that breaks
electron-hole symmetry. For concreteness, we consider the
n-type graphene. In this case the mean free time of excited electrons is
much longer than that of holes since due to different distance from
the Fermi level holes can easier emit phonons. Thus, the contribution of
holes will be neglected.
\begin{figure}
\includegraphics[width=8.5cm]{fig1.eps}
\leavevmode \caption{(Color online) Interband phototransitions in
n-type graphene. Left panel: diagram of transitions in the
momentum-energy space. The hole cone is shifted in $\bf k$ space
by the photon wave vector $\bf q$. The transitions are permitted
only above the Fermi level. Right panel: projection to the
momentum plane. Filled circle represents the Fermi sea,
the elliptic
curve corresponds to the energy conservation equation
$s|{\bf k-q}|+sk=\omega$; only momenta outside the Fermi circle are
permitted corresponding to the right segment of the elliptic curve.
}
\end{figure}
Fig.~2 illustrates a possible experiment on excitation of the drag
current in a suspended graphene sheet placed in $(x,y)$ plane.
Light with frequency $\omega$, wave vector $\bf Q$ ($Q=\omega/c$)
and amplitude of electric field $\mbox{\boldmath{$\cal E$}}$ illuminates graphene
plane. We consider transitions near the cone singularity. In this
case the current is determined by the projections of the electric
field and the wave vector onto the graphene layer.\footnote{The
vertical component of the electric field also interacts with
electrons, however, its action is weaker by the parameter $k_Fd$,
where $d$ is the vertical distance between dangling bonds of
neighboring atoms. In fact, this component results in the
dynamical splitting of these states and can be included in the
Hamiltonian as $\sigma_z eE_z d/2$. Comparison of this term with
considered one gives foregoing estimate.} These quantities are
${\bf E}\equiv{\bf e}E=({\cal E}_p\cos\beta,{\cal E}_s)$ and ${\bf
q}= (1,0)Q\sin\beta$, where $\beta$ is the angle of incidence,
${\cal E}_s$ and ${\cal E}_p$ are amplitude components of the
electric field $\mbox{\boldmath{$\cal E$}}$ perpendicular and parallel to the incident
plane. We ignore small modification of field caused by the layer.
\begin{figure}
\includegraphics[width=9cm]{fig2.eps}
\leavevmode \caption{(Color online)
Sketch of proposed experiment (see text for details).}
\end{figure}
\section{Basic equations}
The current of photon drag effect can be expressed via the
probability of transition $g({\bf k})$ from
the hole state with a momentum ${\bf k-q}$ to the electron state with momentum ${\bf k}$
and the electron velocity ${\bf v}({\bf k})=s{\bf k}/k$ as
\begin{equation}
\label{j}
{\bf j}=4e\int \frac{d{\bf k}}{4\pi^2} {\bf v}({\bf k})\tau g({\bf k}),
\end{equation}
where the coefficient 4 accounts for the valley and spin
degeneracies. The dependence on the photon momentum results from
the momentum and energy conservation laws and the matrix elements
for transition. For simplicity we put below $\hbar=1$.
The two-band Hamiltonian near the Dirac point is
\begin{eqnarray}
\label{1}
\hat{H}({\bf k})=s\left(
\begin{array}{cc}
0 & k_x-ik_y \\
k_x+ik_y&0\\
\end{array}
\right)=s{\bf k}\mbox{\boldmath{$\sigma$}}.
\end{eqnarray}
Here $\mbox{\boldmath{$\sigma$}}$ is the vector of the Pauli matrices.
The eigenvalues and eigenvectors of the Hamiltonian (\ref{1}) are
$\epsilon_{\pm}({\bf k})=\pm sk$ and
$\Psi_{\pm}({\bf k})=(1,\pm e^{i\phi_{\bf k}})/\sqrt{2}$,
where $\phi_{\bf k}$ is the polar angle of the vector ${\bf k}$.
The different signs correspond to
electrons and holes. The interaction with the wave is determined
by the matrix elements of the velocity $\nabla_{\bf k}
\hat{H}({\bf k})=s\mbox{\boldmath{$\sigma$}}$ between the hole and electron states with the
momenta ${\bf k-q}$ and
${\bf k}$: ${\bf v}^{-+}=(\Psi_-({\bf k-q})^*s\mbox{\boldmath{$\sigma$}}\Psi_+({\bf k}))$, correspondingly.
The transition probability $g({\bf k})$ is
\begin{equation}
\label{g} g({\bf k})=\frac{\pi e^2}{2\omega^2}|{\bf E}{\bf
v}^{+-}|^2\delta(sk+s|{\bf k}-{\bf q}|-\omega)\theta(\epsilon_{\bf
k}-\epsilon_F),
\end{equation}
where $\theta(t)$ is the Heaviside function.
The expression for current Eq.(\ref{j}) can be rewritten as
\begin{equation}\label{jj}
{\bf j}=\frac{e^3E^2s^3}{2\pi\omega^2}\int d{\bf k} \tau \frac{\bf k}{|{\bf k}|}
a_{jk}e_je_k\delta(sk+s|{\bf k-q}|-\omega)\theta(sk-\epsilon_F)\end{equation}
where
\begin{equation}\label{a}a_{jk}=
\frac{1}{s^2}v_j^{-+*}v_k^{-+}.
\end{equation}
We utilized the symmetry of the tensor $a_{ij}$ resulting to inclusion of the
field polarization in the combinations $e_i^*e_j+e_j^*e_i$ only
and independence on the degree of circular polarization. Hence,
without loss of generality one can consider the field as
linear-polarized and ${\bf e}$ as real.
Due to the smallness of the wave vector $q$,
as compared to the electron momentum, one can expand all quantities in
powers of $q$.
Expanding by $q$ we can write the argument of the delta-function
as $sk+s|{\bf k}-{\bf q}|-\omega\approx 2sk-\omega -sq \cos\phi_{\bf k}$
(we choose the direction of axis x along ${\bf q}$).
At the same time, $q$ is comparable with $2sk-\omega$
and we keep ourselves from subsequent expansion of the delta-function.
Expanding the tensor $a_{ij}$, we have
\begin{eqnarray}
\label{aa}
a_{xx}&=&\sin^2\phi_{\bf k}(1+\frac{q}{k}\cos\phi_{\bf k}),\nonumber\\
a_{yy}&=&\cos^2\phi_{\bf k}-\frac{q}{k}\sin^2\phi_{\bf k}\cos\phi_{\bf k},\\
2\mbox{Re}(a_{xy})&=&-2\sin\phi_{\bf k}\cos\phi_{\bf k}-\frac{q}{k}\sin\phi_{\bf k}\cos(2\phi_{\bf k})\nonumber
\end{eqnarray}
From Eq.(\ref{jj}) we obtain for components of the current
\begin{eqnarray} \label{jjj}
j_x=-2J_0\int_{-1}^{min(1,a)}
\frac{dx}{\sqrt{1-x^2}}\frac{\tau}{\tau_0}\times\nonumber~~~~~~~~~~~~~~~~~\\
\Big\{e_x^2 [-x(1-x^2)(1+bx)+2b(1-x^2)(1-2x^2)]+ \nonumber \\
e_y^2[-x^3(1+bx)+4bx^2(1-x^2)]\Big\} ;\\
j_y=2J_0e_xe_y\int_{-1}^{min(1,a)} dx\frac{\tau}{\tau_0}\sqrt{1-x^2}\times\nonumber~~~~~~~~~\\
\Big\{-2x(1-x^2)(1+bx)+2b(3x^2-1)\Big\}.
\end{eqnarray} Here we have introduced the following notations:
$$J_0=\frac{e^2}{\hbar c}\frac{cE^2}{8\pi \hbar\omega}|e|\tau_0s, ~~\tau_0=\tau|_{k=k_F} ,~~ a=\frac{\omega-2\epsilon_F}{sq}, ~~ b=\frac{sq}{\omega}.$$
If $\tau$ is independent on the energy of electrons then the integration in
Eq.(\ref{jjj}) can be done directly.
The current has different values inside and outside the region
$|\omega-2\epsilon_F|<sq$. If $|\omega-2\epsilon_F|<sq$ then we have
\begin{eqnarray}
\label{RDE}
j_x&=&-\frac{2}{3}J_0\sqrt{1-a^2}((1-a^2)e_x^2+(2+a^2)e_y^2),\\
j_y&=&-\frac{4}{3}J_0(1-a^2)^{3/2}e_xe_y.
\end{eqnarray}
These values represent resonant photon drag RDE. It remains constant if
$q\to 0$. The value of resonant current is determined by $J_0$.
For the photon flow
$cE^2/8\pi\hbar\omega=10^{19}$cm${}^{-2}$s${}^{-1}$,
$\tau=10^{-12}$ s, $J_0=1.16\cdot 10^{-6}$A/cm. This approximately
corresponds to a power of $0.1W/cm^2$ for photons with energy $0.1eV$.
If $|\omega-2\epsilon_F|>sq$, then there is only NDE current. It is proportional to $q$:
\begin{eqnarray} \label{NDE}
j_x&=&J_0\frac{\pi s q}{4\omega}(3e_x^2-e_y^2),\\
j_y&=&\frac{3}{2}J_0\frac{\pi s q}{\omega}e_xe_y.
\end{eqnarray}
The value of NDE is significantly smaller then the RDE value.
In agreement with the simple estimates the RDE has always the direction
opposite to the direction of light wave vector. Its polarization
dependence is explained by the dependence of the directional
diagram of excitation: most of carriers are excited perpendicular
to the polarization. At the same time the Fermi sea limits the
transitions by the direction of the photon wave vector. This
circumstances together determine lower x-component of current if
${\bf e}||{\bf q}$ in comparison with the case ${\bf e}\perp{\bf
q}$ and also the appearance of y-component of the RDE current.
In agreement with the system symmetry, $j_y$ exists only if the
polarization has both $e_x$ and $e_y$ components. The RDE current
exists in a narrow window $|\omega-2\epsilon_F|<sq$ which shrinks
if $q\to 0$. But inside this window RDE is much stronger than NDE
so the later can be neglected in this window.
\begin{figure}
\includegraphics[width=7.5cm]{fig3.eps}
\leavevmode \caption{ Resonant photon drag current in units of
$J_0$ versus normalized frequency $(\omega-2\epsilon_F)/sq$. The
solid curve shows the longitudinal component of current $j_x$,
the field is polarized along the projection of the wave vector on
the plane ($\theta=0$) and $j_y$ at $\theta=\pi/4$. The dashed
curve shows $j_x$ at $\theta=\pi/2$. }
\end{figure}
\begin{figure}[h]
\includegraphics[width=7.5cm]{fig4.eps}
\leavevmode \caption{Polarization dependence of the RDE current at
$\omega=2\epsilon_F$; $j_x$ is shown by dash curve, $j_y$ is
shown by solid curve.}
\end{figure}
The sign of x-component of NDE depends on polarization. This
contradicts to a simple assumption according to which the current
is mainly determined by kicks which photons give to electrons.
The origin of
this difference is the dependence of the directional diagram on the
small wave vector ${\bf q}$ via the parameter $a_{ij}$: at some
polarizations electrons prefer to be excited in opposite direction
to ${\bf q}$. This explains the change of sign.
Fig.~3 demonstrates the dependence of RDE current components on
the frequency in the window $|\omega-2\epsilon_F|<sq$ where RDE
exists. The current vanishes at the edges of the window. The
component $j_x$ is larger for the polarization along the y axis.
The component $j_y$ appears only for tilted polarization of the
light. Fig.~4 shows the dependence of $j_x$ and $j_y$ on the angle
$\theta$ between the vector of polarization ${\bf e}$ and the
wave vector ${\bf q}$.
\section{Discussion}
We have studied the electron contribution to the photon drag current. In
fact, in the considered system the hole contribution also
presents. The symmetry between holes and electrons in a neutral
system means that these contributions double. However, the result
will be changed if to take into account the difference between
electrons and holes caused by their different excitation energy:
while electrons are generated near the Fermi energy the holes
appear well below the Fermi energy. This leads to a strong
difference between the relaxation times. In high-mobility samples
at low temperature the momentum relaxation time near the Fermi
energy is much greater than far from the Fermi energy. At the same
time, quick relaxation of excited electrons (holes) to the Fermi
energy due to electron-electron interaction (described by e-e
relaxation time $\tau_{ee}$) conserves their momenta up to the
moment when excitations reaches the temperature layer. This
results in equality of holes and electrons contributions to the
current. And vice versa, electron-phonon relaxation can cancel
the hole contribution if $\tau_{e-ph} \ll\tau_{ee}$, where
$\tau_{e-ph}$ is the time of energy relaxation due to
electron-phonon collisions. Thus, the obtained current should be
multiplied by a factor 2 in the case of quick e-e relaxation and
be kept unchanged in the opposite case. We note, that when the Fermi
energy tends to zero the system becomes symmetric.
The RDE exists in a narrow energy range
$\Delta \epsilon \approx \hbar s q \approx \hbar \omega s/c$
near the Fermi energy. This means that the RDE is visible for
temperature $T< \Delta \epsilon$. For photons with $\hbar \omega =0.1 eV $
this gives $T< 3 K$.
The observation of the resonant photon drag in monolayer graphene
is accessible to the modern experimental technique that allows to
investigate interesting aspects of coupling between photons and
electrons in this material.
\section{Acknowledgments}
We thank A.D.Chepelianskii for useful discussions.
The work was supported by grant of RFBR No 08-02-00506 and No
08-02-00152 and ANR France PNANO grant NANOTERRA;
MVE and LIM thank Laboratoire de Physique Th\'eorique, CNRS
for hospitality during the period of this work.
|
\section{Introduction}
Symmetrical morphologies and regular patterns in living organisms (Fig~\ref{Fig1}) have been credited with originating the idea of beauty, the notion of art as an imitation of nature, and humanity's first mathematical inquiries~\cite{Adler,Jean,Smith,Grew}. The fascinating symmetrical patterns of organs in plants, called phyllotaxis~\cite{Adler,Jean, Smith}, were known to the Romans (Pliny) and ancient Greeks (Theofrastus), while early recognitions are found in sources as ancient as the Text of the Pyramids~\cite{Adler}. Leonardo da Vinci~\cite{daVinci}, Andrea Cesalpino, and Charles Bonnet~\cite{Bonnet} studied phyllotaxis in the modern era. Kepler proposed that the Fibonacci sequence (1, 2, 3, 5, 8\dots), where each term is the sum of the two preceeding ones~\cite{Fibonacci}, describes these phyllotactic patterns.
A discipline that thrived on multidisciplinary interactions~\cite{multi}, phyllotaxis found its standard mathematical description when August and Louis Bravais~\cite{Bravais} introduced the point lattice on a cylinder to represent the dispositions of leaves in 1837 (see Fig.~\ref{Lattice}), thirteen years {\it before} August's seminal work on crystallography~\cite{Bravais2}. Unfortunately botanists neglected the work of the Bravais brothers, and it wasn't until Church rediscovered it eighty years later that more progress was achieved in the field~\cite{Church}.
The geometrical description of cylindrical phyllotaxis relies, in the simplest case, on the phyllotactic lattice introduced by the Bravais brothers~\cite{Adler,Jean,Smith,Bravais}. It consists of a so-called generative spiral of divergence angle $\Omega$. We can visually decompose the resulting lattice in crossing spirals that join nearest neighbors, as in Fig.~\ref{Lattice}, which botanists call
parastichies. It is a fundamental observation (made first by Kepler) that the numbers $n$, $m$ of crossing parastichies needed to cover the lattice are consecutive terms of the standard Fibonacci sequence, or less frequently the variants obtained by changing the second term, also called Lucas numbers: 1, 3, 4, 7, 11\dots and 1, 4, 5, 9\dots often referred to as second and third phyllotaxis. From that, one can prove that the divergence angle of the generative spirals in plants assumes values close to~\cite{Jean,Adler74}
\begin{equation}
\Omega_p=\frac{360 ^\circ}{\left(\tau +p\right)},
\label{Omega}
\end{equation}
where $p=1, 2, 3$ denotes first, second or third phyllotaxis and $\tau=\left(1+\sqrt{5}\right)/2$ is the golden ratio. For more than one generative spiral (``multijugate'' phyllotaxis), parastichies share a common divisor $(n,m)=(k n',k m')$, $k$ being the number of generative spirals~\cite{Adler74,Jean}. Not unlike domain boundaries in crystals, plants show kinks between domains, called transitions by botanists~\cite{kink,Adler}.
In the last 50 years, phyllotactic patterns have been seen or predicted outside of botany: polypeptide chains~\cite{Frey, Abdulnur}, tubular packings of spheres~\cite{Erickson}, convection cells~\cite{Rivier}, layered superconductors~\cite{Levitov}, self-assembled microstructures~\cite{Chaorong}, and cooled particle beams~\cite{Rahman,Shatz}. While it is still debated whether such systems might shed light on botanical phyllotaxis, the occurrence of such mathematical regularities outside of botany is fascinating and leads to generalizations that -- unlike quasistatic botany -- allows for dynamics~\cite{Nisoli_PRL}.
In a groundbreaking work Levitov recognized phyllotaxis in vortices of layered superconductors~\cite{Levitov}. He next described how phyllotactic patterns represent states of minimal energy of a cylindrical lattice (that is of a lattice with cylindrical boundary conditions) of mutually repelling objects, the repulsion mimicking the interactions between spines, leaves, or seeds in plant morphology~\cite{Levitov2, Levitov3}. Yet such a constraint to a lattice is absent both in botany and in the physical systems to which this energetic model might apply, such as adatoms or low-density electrons on nanotubes and ions or dipolar molecules in cylindrical traps.
\begin{figure}[t!]
\begin{center}
\includegraphics[width=3 in]{Fig1_s.jpg}
\caption{Natural and Magnetic Cacti. A specimen of {\it Mammillaria elongata} displaying a helical morphology ubiquitous to nature, a magnetic cactus of dipoles on stacked bearings, and a schematic of a wrapped Bravais lattice showing the angular offset (divergence angle) $\Omega$ and the axial separation $a$ between particles.}
\label{Fig1}
\end{center}
\end{figure}
Following up on earlier work that focused on the dynamics of rotons and solitons in physical phyllotactic systems~\cite{Nisoli_PRL}, we provide here a detailed experimental and numerical demonstration that Levitov's constraint is not necessary, and that the lowest energy states of repulsive particles in cylindrical geometries are indeed phyllotactic lattices. In addition, we describe the experimental and numerical generation of multijugate phyllotaxis, static kink-like domain boundaries between different phyllotactic lattices, and unusual disordered yet reflection-symmetric structures that may be a static relic of soliton propagation.
We show that when a ``magnetic cactus'' of magnets (spines) equally spaced on co-axial bearings (stem) with south poles all pointing outward is annealed, it precisely reproduces botanical phyllotaxis. When studied numerically via a structural genetic algorithm, the fully unconstrained case reveals both multijugate and mono jugate phyllotaxis. In addition to our macro-scale implementation, such systems could also be created at the quantum level in nanotubes or cold atomic gases.
In section II we describe the statics of repulsive particles in cylindrical geometries. In section III we detail the experiment on the magnetic cactus. In Section IV we discuss the more general case of multijugate phyllotaxis.
\begin{figure}[t!]
\begin{center}
\includegraphics[width=2.8 in]{Bravais_01}
\caption{The Bravais lattice with cylindrical boundary conditions that defines a phyllotactic spiral. The cylinder axis is vertical, while the horizontal direction contains three circumferential repeats. The solid line is the generative spiral: this one-dimensional Bravais lattice generates the full structure. The dashed lines are the so-called parastichies or visible secondary spirals: they connect nearest neighbors on the surface of the cylinder. Adapted from A. Bravais and L. Bravais, 1837~\cite{Bravais}.}
\label{Lattice}
\end{center}
\end{figure}
\section{Phyllotaxis of repulsive particles in cylindrical geometries}
In this section we will recall Levitov's model~\cite{Levitov2, Levitov3} and some of our own findings~\cite{Nisoli_PRL}. Following Levitov, let us assume that the lowest energy configuration for a set of particles with repulsive interactions, confined to a cylindrical shell of radius $R$, is a helix with a fixed angular offset $\Omega$ between consecutive particles and a uniform axial spacing $a$, as in Fig.~\ref{Fig1} (this so far unproved ansatz will be investigated later both numerically and experimentally). For a generic pair-wise repulsive interaction $v_{ij}$ between particles $i$ and $j$, the energy of the helix is $V = \frac{1}{2}\sum_{i \neq j} v_{i,j}$. Since the lattice structure is defined by $\Omega$, we can write $V(\Omega)$.
In Fig.~\ref{spectra} we plot $V\left(\Omega\right)$ for various values of the ratio $a/R$:
for specificity we employed a dipole dipole interaction
$v_{i,j}={\bm p}_i \cdot {\bm p}_j/ r_{i,j}^3 -3 ({\bm p}_i \cdot {\bm r_{i,j}})({\bm p}_j\cdot {\bm r_{i,j}}) /r_{i,j} ^5 $, repulsive at the densities considered here. However, the following considerations only depend upon geometry and therefore apply to a vast range of
reasonably behaved, long range repulsive interactions.
When $a/R \gg 1$, the angle $\Omega = \pi$ maximizes distance between neighboring particles and therefore $V\left(\Omega \right)$ has a minimum in $\pi$. The angle between second nearest neighbors along the helix is $2\pi$, which means that they face each other.
And thus, as the density increases, interaction between the facing second nearest neighbors becomes predominant, and $\Omega = \pi$ is not a minimum for $V\left(\Omega \right)$ anymore. If whe shift the helical angle from $\pi$, the repulsive interaction between second nearest neighbors is reduced, with minimal penalty from nearest neighbors. In terms of $V\left(\Omega \right)$, that means a local maximum $\Omega=\pi$.
This argument can be iterated for every commensurate winding that allows particles separated by $j$ neighbors to face each other. As density increases further, the angles $2\pi/3$ and $4\pi/3$ also become unfavorable due to third-neighbor interactions. Any commensurate spiral of divergence angle $\Omega=2\pi i/j$ with $i, j$ relatively prime corresponds to a configuration where each particle faces each $jth$ neighbor. For every $j$ there will be a value of $a/R$ low enough such that $\Omega=2\pi i/j$ is a local maximum, which we call a peak of rank $j$.
The proliferation of peaks for increasing linear density is shown in Fig~\ref{spectra}. We can see that at high density, peaks of equal rank are nearly degenerate; that is natural, since their principal defining energetic contribution arises from particles facing each other at a distance $j a$. The minima also become more nearly degenerate as the density increases. That can be explained intuitively, since for angles incommensurate to $\pi$ each particle ``sees'' the others as incommensurately smeared around the cylinder, and is therefore embedded in a nearly uniform background charge from the other particles. The degenerate energy of the ground state can be well approximated by an uniform continuum distribution $\epsilon_0$, whereas the energy of a peak of order $j$ will be $V\left(2\pi i/j\right)\simeq v(j a)+\epsilon_0$, where $v(r)$ is the energy of two particles facing at a distance $r$: for our dipole interaction $v(ja)=p^2/a^3 j^3$.
The first step to calculate the degeneracy of our system at a given density, is to find the corresponding maximum rank of the peaks. As all of the peaks of the same rank have the same energy, and appear in the spectrum together, we can focus on the emergence of $2\pi /J$. For $a/R\ll1$, this new peak will emerge when the distance between particles separated by a distance $Ja$ equals that of particles separated by $2\pi R/J$. Therefore one finds for the maximum rank
\begin{equation}
J = \Big\lbrack{\kern -0.1 em}\Big\lbrack
\sqrt{\frac{2\pi R}{a}} \Big\rbrack{\kern -0.12 em}\Big\rbrack,
\label{J}
\end{equation}
which as expected only depends on purely geometrical parameters. A little
more tricky is to compute the degeneracy, given $J$. The set of all the peaks has the cardinality of the class of all the fractions $i/j$, with $i,~j$ coprime and $j\le J$. This can be considered as the disjointed union of other classes, called Farey classes of order j, defined as follows: $P_j\equiv \{\Omega = 2\pi i/j\mid$ for $i, \!\ j$ coprime and $i\leq j\}$, i.e. all fractions in lowest terms between 0 and 1 whose denominators do not exceed $j$~\cite{Farey}. The union of all Farey classes up to a certain order $J$ has the cardinality of the set of peaks for a spectrum of maximum rank $J$. Now, the cardinality of $P_j$ is know from number theory to be Euler's totient
function, $\phi (j)$~\cite{Totient}. Therefore, the degeneracy $D$ of the energy minima for a system with a maximum peak rank $J$ is~\cite{Totient}
\begin{equation}
D=\sum_{j=1}^{J}\phi(j)=\frac{3}{\pi^2} J^2 +O\left(J \log J\right),
\end{equation}
which, from Eq.~\ref{J}, scales as $D\sim 2R/a$.
Finally, we recall~\cite{Smith,Levitov} that the order $j_1$, $j_2$ of the peaks bracketing a minimum relates to its structure in a straighforward way: the helix corresponds to a rhombic lattice where each particle has its nearest neighbors at axial displacements of $\pm a j_1$, $\pm a j_2$ and second nearest neighbors at $\pm a (j_1+j_2)$ or $\pm a (j_1-j_2)$~\cite{Levitov}. Also, $j_1$ and $j_2$ give the number of crossing secondary spirals (parastichies) needed to cover the lattice by connecting nearest neighbors.
\begin{figure}[t!!!!!!!!!!!!!!!]
\center
\includegraphics[width=3.2 in]{fig2}
\vspace{-2mm}
\caption{Lattice energy $V(\Omega)$ versus divergence angle for successively halving values of $a/R$ starting from $0.5$ (using dipole dipole interaction, $\epsilon=p^2/a^2$, where $p$ is the magnetic dipole). Notice the proliferation of peaks as $a/R$ decreases. Reproduced from \cite{Nisoli_PRL}.}
\label{spectra}
\vspace{-2mm}
\end{figure}
For completeness, let us now follow Levitov~\cite{Levitov,Levitov2,Levitov3}, and consider the adiabatic evolution of our system as the linear density is increased. As new sets of maxima and minima emerge,
the true minimum goes through a series of quasi-bifurcations, the consequence of an elusive symmetry whose explanation goes beyond our scope.
Suffice it to say that the system evolves quasi-statically from one of these optimal $\Omega$ to another as $R/a$ increases, asymptoting to the golden angle $\Omega_1=2\pi/\left(\tau +1\right)$ $\left[\tau=\left(1+\sqrt{5}\right)/2\right]$, ubiquitous in botany, as each minimum is bracketed by peaks whose ranks, because of the Farey tree structure described above, are consecutive elements of the Fibonacci sequence. Occasional ``wrong turns'' at later stages, will not shift the convergence too far from the golden angle, yet the Fibonacci structure would be lost. However if one or two consecutive wrong turns happen at the second or second and third bifurcations the system will converge to the alternative angles of second or third phyllotaxis, given by Eq.~(\ref{Omega}).
We have only surveyed so far spiraling lattices generated by a single helix. A straightforward generalization gives multijugate phyllotaxis, when two or more elements grow at the same axial coordinate~\cite{Adler, Jean, Smith}. This case, which Levitov does not explore, can be easily mathematically reduced to monojugate case, by considering two or more replicas of the phyllotactic lattice as in Fig.~\ref{Lattice}. In our experimental realization we restrict ourselves to the monojugate phyllotaxis, and explore multijugate only numerically.
\section{A Magnetic Cactus}
There is a long history of experimental reproductions of phyllotactic patterns. Recently, Doady {\it et al.} described phyllotaxis in terms of dynamical systems and then verified it experimentally by examining dynamical processes in droplets of ferro-fluid~\cite{Douady}. But even more than a century ago, Airy showed that phyllotaxis emerged in optimal packing of hard spheres connected by a rubber band, once the band was twisted to increase density~\cite{Airy}.
Here we expand on what was announced in a recently published Letter~\cite{Nisoli_PRL}: we verify experimentally the assumptions of Levitov's energetic model, by studying the low energy configurations of interacting magnets stacked evenly-spaced and free to rotate around a common axis. We constructed a mechanical system that it is free to explore the three angles of botanical phyllotaxis (Eq.~\ref{Omega}).
\subsection{Experimental Apparatus}
\begin{figure}[t!!!!!!]
\center
\vspace{3 mm}\includegraphics[width=2.7 in]{exp_app_2}\vspace{10 mm}
\includegraphics[width=2.9 in]{exp_app_1}
\caption{Experimental apparatus. Top: Each unit of the magnetic cactus consists of a magnet element and a unit ring secured to a central axis. The ring diameter \emph{d} is 2.2 cm. Bottom: a schematic representation of the mounted
magnetic cactus and surrounding measurement devices. The viewer's eye is restricted by the viewing slit and the reference wires. Measurements are taken
directly from the dividing head.}\vspace{-12mm}
\label{apparatus}
\end{figure}
We built a magnetic cactus by mounting permanent magnets (spines) on stacked co-axial bearings (a stem) which are free to rotate about a central axis, as in Fig.~\ref{apparatus}. All the magnets point outward, to produce a repulsive interaction between all magnet pairs. To avoid effects of gravity, the apparatus rests in the vertical position, and is non-magnetic. We built two different versions, the second with magnets twice as long as the first, as to have a larger effective radius which gives three rather than two stable structures.
We employed cylindrical permanent iron-neodymium magnets, 1.2 cm long and 0.6 cm in diameter. They are mounted on fifty aluminum rings of 2.2 cm outer diameter, each affixed to a non-magnetic radial ball bearing (acetal/silicon, Nordex) as in Fig.~\ref{apparatus}. These unit rings are evenly spaced on an aluminum rod in a stacked structure of 39.9 cm axial length.
At static equilibrium we measure separation angle between each magnet element, by rotating the cactus until a magnet element aligns with the reference wires.
A telescope and a vertical viewing slit accompanied by two vertical reference wires assist in data acquisition.
\begin{figure}[t!]
\center
\hspace{10 mm}\includegraphics[width=2.2 in]{Fig2_a}
\hspace{10 mm}\includegraphics[width=2.2 in]{Fig2_b}\vspace{3 mm}
\includegraphics[width=2.9 in]{onemagnet3}\vspace{-.5 mm}
\includegraphics[width=2.9 in]{twomagnet3}
\caption{Top: A 3-D rendering of the experimental data and the corresponding Bravais lattice of the magnetic cactus annealed in a spiral configuration of divergence angle $\Omega_1$ and parastichies (2,3) (blue and red dashed lines), Fibonacci numbers. The bottom two panels show the experimentally measured angular offsets $\Omega$ between successive magnets for magnetic cacti with short (middle) and long (bottom) magnets, plotted versus magnet index, which simply counts the number of magnets along the axis. Flat regions are perfect spirals while steps are boundaries between different phyllotactic domains. The dotted lines give the phyllotactic angles $\Omega_1$, $\Omega_2$, $\Omega_3$ and $2\pi-\Omega_2$ defined in the text, whereas the dashed lines are minima of the magnetic lattice energy (insets) calculated by interpolating the measured pair-wise magnet-magnet interaction. Data reproduced from Ref. \cite{Nisoli_PRL}.}
\label{Fig2}
\end{figure}
\subsection{Annealing}
By measuring the dipole-dipole interaction between an individual magnet pair, we can reconstruct the curve of the lattice energy $V(\Omega)$ as a function of the angular offset $\Omega$ between magnets. We find that the first arrangement, with short magnets, admits two minima, given by the angles of Eq.~\ref{Omega} for $p=1,2$. The second arrangement, with long magnets, has three minima corresponding to the angles of Eq.~\ref{Omega} $p=1,2,3$, one of which ($p=3$) is a weak metastable minimum. These divergence angles of stable helices are very close to those predicted by phyllotaxis, of Eq.~\ref{Omega}, and are all accessible by experimental procedure described below.
Before every data acquisition, the cactus is disordered and then athermally annealed into a low-energy state. The protocol involves repeatedly winding the bottom-most magnet to generate an ever-tightening spiral, until an explosive release of energy disorders the lattice. Next, an independent external magnet is oscillated in small circular motions near randomly chosen points while the cylinder as a whole is slowly rotated, to further randomize magnet orientations. After 10-30 second of mechanical annealing through applied vibrations, the system consistently enters a robust ordered state which does not anneal further on experimental timescales.
\begin{figure}[t]
\hspace{14 mm}\includegraphics[width=2.3 in]{Fig3_a}\vspace{5 mm}
\hspace{15 mm}\includegraphics[width=2.2 in]{Fig3_b}\vspace{5 mm}
\includegraphics[width=2.9 in]{Fig3_c}\vspace{5 mm}
\caption{ A kink between domains of first and second phyllotaxis. From top to bottom: a 3-D rendering, and its unwound 2-D Bravais lattice from numerical simulations. Below, the same kink plotted as angle increments between successive magnets for the experimental system (crosses) and numerical simulation (circles).}
\label{kink}
\end{figure}
\subsection{Results}
Figure~\ref{Fig2} reports the experimental results for both arrangements by plotting the measured angle between consecutive magnets. The more narrow (short-magnet) cactus self-organizes into the spirals with divergence angles precisely reproducing those of first phyllotaxis, $\Omega=\Omega_1$, and second phyllotaxis, $\Omega=\Omega_2$, as in Eq.~(\ref{Omega}). When the results are represented in a 2-D lattice, as in the top of Fig.~\ref{Fig2}, parastichies can be drawn. As parastichial numbers for $\Omega=\Omega_1$ we find the Fibonacci numbers $(2,3)$, and for $\Omega=\Omega_2$, the Lucas numbers $(3,4)$, as seen also in botany.
The larger-radius system also forms first and second Phyllotaxis helices, as well as limited domains of third phyllotaxis [with $\Omega=\Omega_3$ and Lucas numbers $(1,4)$], bracketed by domains of second phyllotaxis. The
insets of Fig.~\ref{Fig2} show the magnetic interaction energy $V\left(\Omega\right)$ of the lattice obtained by interpolating measured values for the pair-wise magnet-magnet interaction, plotted as a function of divergence angle $\Omega$. As we cans see, local minima correspond to phyllotactic angles.
\begin{figure}[t!!]
\begin{center}
\hspace{13 mm}\includegraphics[width=2.5 in]{fig7a}\vspace{4mm}
\includegraphics[width=3.2 in]{fig7b}
\hspace{12 mm}\includegraphics[width=3.2 in]{fig7}
\end{center}
\caption{A numerically calculated kink in a system of dipoles of high degeneracy (seven minima, $a/R=0.15$). The kink separates domains with parastichy numbers (4,5) and (5,6) and divergence angles of 1.38 and 1.13 radians. The top and middle panels give its three-dimensional rendering and angular shift $\Omega$ versus the axial magnet index. The two domains correspond to minima bracketed by peaks at $\Omega/2\pi = 1/4, 1/5$ and $\Omega/2\pi = 1/5, 1/6$ of the lattice energy, given in the bottom panel, where $\epsilon=p^2/a^2$, $p$ being the magnetic dipole.}
\label{kink2}
\vspace{-2mm}
\end{figure}
\begin{figure}[t!]
\center
\includegraphics[width=3. in]{phantom}
\caption{Measured angular offsets between successive magnets for magnetic cacti with long magnets, plotted versus magnet index, showing symmetric kink/anti-kink domain boundaries. Dotted lines give the phyllotactic angles $\Omega_1$, $\Omega_2$, $\Omega_3$ and $2\pi-\Omega_2$ defined in the text. Dashed lines are minima of the magnetic lattice energy (inset) calculated by interpolating the measured pair-wise magnet-magnet interaction.}
\label{phantom}
\end{figure}
Figure~\ref{Fig2} also shows that in many instances the system fragments into two or three distinct domains whose domain walls always share a common parastichy, as seen in botany~\cite{kink,Adler}, and as expected in physics for a quasi-one-dimensional degenerate system. We have computed numerically one such transition via dynamical simulations in a velocity-Verlet algorithm, in the following way: we start from a crude static step-like kink as an initial condition, and allow it to radiate energy in the form of phonons waves until it stabilizes in a kink with superimposed vibrations; we then average this configuration over time, to remove these residual oscillations. When the result is used as new initial conditions, it proves to be a static kink. Fig.~\ref{kink} reports our numerical results for a kink in a system whose size and interaction reproduces the physical realization of the magnetic cactus, along with the experimental data for such a kink. The match is essentially perfect, indicating that the dissipative (i.e. frictional) forces neglected in our model do not significantly affect the static configurations. We apply the same numerical procedure to calculate a kink in a system of larger degeneracy, among domains which are absent in our physical realization. We use a smaller $a/R$ ratio and a different interaction between particles (ideal dipole instead of physical dipole). The result shown in Fig.~\ref{kink2} reproduces the same qualitative shape of the previous, lower degeneracy case. Similar kinks are present also in a fully unconstrained cactus, one in which the particles can move along the axis, and are found in early generations of our structural genetic algoritms (see below).
Finally, these kinks can travel as novel topological solitons, with a rich phenomenology that is explained elsewhere~\cite{Nisoli_PRL, Nisoli_PRE}.
Finally, in the system with longer magnets, we occasionally found intriguing yet hard-to-interpret configurations that contain two nearly reflection symmetric domain boundaries. Fig.~\ref{phantom} reports two such configurations, measured in independent experimental runs. Although we do not have a firm explanation for these structures, we speculate that they form as frozen-in soliton waves that initiated symmetrically at both ends of the structure, upon release of the wound-up elastic energy during initial preparation. Indeed an analytical, continuum theory for phyllotactic solitons which we have developed recently, and which explains results of the dynamical symulations also supports the existence of similar frozen-in pulses~\cite{Nisoli_PRE}.
\section{Fully unconstrained cactus: structural genetic algorithm}
Our experimental apparatus is not fully unconstrained: the axial coordinates of the dipoles are fixed, and so only the azimuthal movement is allowed. While this is an huge improvement toward the original helical constraint of Levitov, many (most) physical systems that could manifest phyllotactic patterns do not posses such a lesser azimuthal constraint. To corroborate and extend our experimental results to a completely unconstrained system, we seek the energy minimum in a set of repulsive particles that can move {\it axially} as well as {\it angularly} on a cylindrical surface, via a non-local numerical optimization. To this purpose, we developed a structural genetic algorithm.
\subsection{Genetic Algorithm}
A genetic algorithm is a method of optimization that mimics evolution to find the absolute minimum in a function which shows a large number of metastable minima. The coordinates of the energy functions are called genes, and a set of genes is a particular specification of value for those variables. The routine typically starts with a set of ``parents'', or specific points in the domain of the energy function. At each stage of the routine, parents ``mate'' to produce children via exchange of genes: a subset of the coordinates of each of the two configurations are swapped, therefore generating new points in the energy domain, called children. Each of those children is then locally relaxed to a minimum via a local search. The new population of parents and children undergoes genetic selection and only the fittest (the lowest energy ones) form a new population.
There are many different implementations of this general idea: care is taken not to lose genetic diversity during selection, to avoid a population of almost identical replicas; that is usually achieved with a more or less skilled genetic selection, which might retain less geneticaly fit individuals, and often by introducing mutations in the form of random alteration of the gene sequence, which would opefully prevent the routine from gettting stuck around a metastable region. Choice of parameterization of the structures (genes) and mating (crossover) is crucial to the performance of the algorithm.
About fifteen years ago, Deaven and Ho~\cite{Deaven} introduced a so-called {\it structural} genetic algorithm, which proved particularly efficient in minimazing the energy of physical structures, as it allows for physical intuition in defining the genes and mating procedure. With it, they found the C$_{60}$ fullerene structure as a ground state of 60 carbon atoms interacting with suitable atomic potentials~\cite{Deaven} and solved the celebrated Thomson problem of repulsive charges on a sphere~\cite{Morris}, a task quite similar to ours.
\begin{figure}[t!!!]
\begin{center}
\hspace{14 mm}\includegraphics[width=3.1 in]{energy.pdf}
\caption{Relative energy of the fittest member of the population in every generation. Early generations return very high energy configurations corresponding to disordered metastable states. Interrmediate generations show populations of phyllotactic domains separated by kinks between. Finally, the algorithm converges to a single crystalline domain in the bulk (deformations at the boundaries simply accomodate the system to the confining potential).}
\label{GAenergy}
\end{center}
\end{figure}
\subsection{Our Algorithm}
In our implementation we use a population of 10 members. Each member reppresents a configuration of 101 particles on the cylinder: more esplicitly, the genetic structure of each member $P_k$, $k=1,\dots,10$, of the population is a set of variables, or $P_k=\{ \theta_i,z_i \}^{101}_{i=1}$ which specifies, in cylindrical coordinates, the positions of the particles composing its structure. The particles interact via a pair-wise inverse quadratic repulsion $V=V_o (r_o/r)^2$, where $r$ {\it is the three dimensional distance between particles}; we introduce a confining potential in the form of an external axial square-well of width $L$, which sets the length of the cylinder, and hence the density. The choice of the pairwise interaction is not fundamental, as long as it is long ranged, repulsive, and well behaved~\cite{Nisoli_PRL}; our particular choice simply speeds up the computation.
We generate the first population randomly. At each step, we randomly couple mates, and excahge their genes employing the following mating procedure: we order the genes by increasing axial coordinates $z_1<z_2<\dots<z_{101}$ and swap the first $1<n<101$ genes, where $n$ is a random number, between randomly selected parents. The children obtained in this way are then relaxed to a stable structure via a standard conjugate gradient algorithm. We then prepare the new generation by selecting the lowest energy individuals in the population of parents and children, yet making sure that the energy difference between members does not fall below a certain threshold, to preserve genetic diversity: when new children cannot produce a new population of 10 in accordance with the energy threshold, we introduce mutations by randomly altering a certain number of members.
\begin{figure}[t!!]
\begin{center}
\hspace{14 mm}\includegraphics[width=2.27 in]{Fig8a}\vspace{5mm}
\includegraphics[width=2.9 in]{Fig8b}
\caption{Numerical optimization via structural genetic algorithm for $N=101$ repulsive particles [$V=V_o (r_o/r)^2$] constrained to a cylindrical surface of length $L$ and radius $R=1.65 L/N$. The resulting 2-D Bravais lattice has a nearly constant axial separation $\Delta z=z_{i+1}-z_i$ (top) and angular divergence $\Omega$ between successive particles (bottom), neglecting fringe effects at the border of the potential well. In the bulk, particles self-organize on a single spiral of divergence $\Omega=\Omega_1$. Oscillations at the boundaries are due to the effect of the confining potential.}
\label{Fig4}
\end{center}
\end{figure}
\subsection{Numerical Results}
During the structural evolution, the earliest populations contains metastable disordered states. Members of intermediate populations show kinks between domains of different divergence angle, configurations which are also seen experimentally. After fifteen to twenty generations, the algorithm typically converges to a single crystalline domain.
Figure~\ref{GAenergy} reports the energy of the fittest member of the population at each generation in a typical run, showing a punctuated-equilibrium evolution where the most-fit structure progressively decreases in energy in intermittent steps separated by plateaux. The final converged results form well-defined two-dimensional cylindrical crystals away from boundaries.
Figure~\ref{Fig4} shows the crystalline structure to which the algorithm converges, for $R=1.65 L/N$: a single spiral with $\Omega = \Omega_1$, as defined in Eqn.~(\ref{Omega}), corresponding to first phyllotaxis with parastichies $(1,2)$. A plot of $\Delta z=z_{i+1}-z_i$ returns the value $L/N$ in the bulk, which implies a single generative spiral. This choice of $R N/L$ corresponds to a density close to that of our experimental apparatus.
\section{Multijugate phyllotaxis}
\begin{figure}[t]
\begin{center}
\hspace{7 mm}\includegraphics[width=2.8 in]{p2}\vspace{6 mm}
\hspace{-4.55 mm}\includegraphics[width=3.15 in]{multi}
\caption{ Top: 3-D schematics and 2-D lattice for a 2-jugate configuration. Bottom: Lattice energy versus angular offset for n-jugate configurations in a system of repulsive dipoles ($\epsilon=p^2/a^2$, where $p$ is the magnetic dipole) for 1-jugate (black), 2-jugate (red), 3-jugate (green), and 4-jugate (blue).}
\label{multi}
\end{center}
\end{figure}
For highly degenerate systems the genetic algorithm returns configurations with more than one generative spiral, corresponding to what in botany is called multijugate phyllotaxis~\cite{Jean}. We have seen before that helices make cylindrically symmetric lattices. On the other hand, every cylindrically symmetric lattice can be decomposed into a suitable number of equispaced generative spirals~\cite{Jean, Bravais}. That is accomplished by discretizing the cylinder along its axis into equally spaced rings and then assigning at each ring $n$ sites, equally spaced and separated by a $2\pi/n$ angular shift. As before, each ring is shifted consecutively by a divergence angle $\Omega$. The case $n=2$ is shown at the top of Fig.~\ref{multi}. The case $n=1$ is shown in our experimental arrangement of Fig.~\ref{Fig1}.
By decomposing the n-jugate cylindrical lattice into $n$ lateral replicas of single-spiral lattice, as in Fig.~\ref{Lattice}, the reader is easily convinced that multijugate phyllotaxis reduces to the previously described monojugate case. All the considerations above apply, provided that one now takes the periodicity to be $2\pi/n$, and the distance between rings to be $n a$ (with, as before, $a=L/N$). It follows that a n-jugate configuration will have local maxima in $2\pi i/n$, $i=1\dots n$ and, following the discussion of section II, one finds that there will be other local maxima corresponding to angles $2\pi/n\times i/j$ when $j \le [[J/n]]$, and $J$ is the maximum rank given by Eq.~\ref{J}.
Note now that if two multijugate lattices of jugation $n$, $n'$ have a peak in the commensurate angle $2\pi i/j$, then the energy of the peak is the same, as is shown in Fig.~\ref{multi}, bottom, which compares the plots of the energy of such an arrangement for different values of $n$. In fact both configurations correspond to particles facing each other after $j/n$ and $j/n'$ rings, and therefore at the same distance $na\times j/n= n'a\times j/n' =j a$, independent of $n$ or $n'$. For small $n$ the minima in the energy of n-jugate configurations essentially degenerate with the monojugate one previously explored. For large $n$ they have higher energy. If $a/R$ is small enough, the threshold is $n>J$, as interaction between particles on the same ring become comparable to those facing in the minimal monojugate peak.
\section{Conclusion}
We have studied the lowest energy configurations of repulsive particles on cylindrical surfaces, both experimentally and numerically. We have found that they correspond to the spiraling lattices seen in the phyllotaxis of living beings, both monojugate and multijugate. By establishing experimentally and numerically that phyllotactic point lattices are ground states in the very general geometric scenario of unconstrained repulsive particles on cylinders, we have opened the study of phyllotaxis to a much wider range of annealable physical systems where the particles could be electrons, adatoms, ions, dipolar molecules, nanoparticles, etc. constrained by external potentials.
Unlike plants, these multifarious, non-biological Phyllotactic systems could access various degrees of dynamics, providing new phenomenology well beyond that available to over-damped, adiabatic botany. We have reported elsewhere~\cite{Nisoli_PRL} on the dynamical richness of this physical phyllotaxis, including classical rotons and a large family of novel, inter-converting topological solitons.
|
\section{Introduction}
\label{int}
\vskip .1cm
The story of the zero-field susceptibility $\chi$ of the two-dimensional Ising model is a landmark
saga of mathematical physics. A recent review of the highlights can be found in \cite{M10}.
While a closed form expression for the susceptibility still eludes us, we possess an
enormous amount of exact or extremely precise numerical information. This largely
derives from two complementary approaches. One approach involves studying the
series expansion of the susceptibility. Since the work of Orrick \etal \cite{ONGP01}
we have had available a polynomial time algorithm, now of complexity O$(N^4)$
for a series of length $N$ terms. From the point of view of an algebraic combinatorialist,
this comprises a solution, and many questions about the asymptotics, and about the
scaling functions have been answered by analysis of the very long series we now
have available--currently more than two thousand terms in length.
The difficulty in proceeding further with this approach is that we have no idea what the
underlying closed-form solution looks like, except that it is known, or more precisely
universally believed, to be non-holonomic. The alternative approach is to express the
susceptibility as a form-factor expansion. This approach was initiated more than 30
years ago by Wu \etal \cite{WMTB76}. In this representation, the susceptibility is written
\begin{equation}
k_BT\, \chi = (1-t)^{1/4}\sum_{n \ge 1} {\tilde \chi}^{(2n+1)}
\end{equation}
for $T > T_c,$ where $t = \sinh^4 (J/k_B T).$
For $T < T_c$ a similar expression with even superscripts prevails. The advantage of
the form-factor approach is that each term in the sum {\it is} holonomic. This means that,
with sufficient computational resources, and sufficient ingenuity, each term can be found.
To date, the first six terms have been found, in the sense that their defining ODE has
been obtained, either totally, or {\it modulo} a prime. We also have precise integral
representations for the form-factor terms, and a Landau analysis of the integrands can
provide information as to the distribution of singularities in the complex plane. Indeed,
it was just such a study by Nickel \cite{nickel-99, nickel-00} that gave convincing evidence
of a natural boundary in the {\it total} susceptibility, thus supporting an earlier but weaker
argument of Guttmann and Enting \cite{GE96} that the total susceptibility was non-holonomic.
While an exact solution for the Ising model susceptibility may be impossible and is certainly
beyond reach at present, one might hope to obtain a complete picture of the singularities of
the susceptibility. Indeed, this more limited goal has been the main motivation of the recent
studies of the individual form-factor terms. For this the ODE and Landau analysis approaches
are complementary; the Landau analysis provides necessary but not sufficient conditions \cite{Hwa}
while the ODE, even if only in mod prime representation, can show which Landau singularities
are to be excluded. The most detailed study in this regard is that of the five particle contribution
$\tilde{\chi}^{(5)}$ to the susceptibility initiated by Boukraa \etal \cite{experim} and followed by
Bostan \etal \cite{High}. The present paper is an attempt to address a number of issues
left unresolved in these papers.
A brief summary of the parts of \cite{experim} and \cite{High} relevant here is as follows.
In Boukraa \etal \cite{experim} series in $\, w$ modulo a prime to 10000 terms were given
and shown to be adequate to find the order 33 Fuchsian differential
equation\footnote[2]{The notation used here is that in~\cite{experim}: variables
$\, w$ and $\, s$ are useful for high temperature expansions with $\, w\, =\, s/2/(1+s^2)$.
The high temperature $\, \chi^{(2n+1)}$ and $\, \tilde{\chi}^{(2n+1)}$ are related by
$ \, s \cdot \chi^{(2n+1)}\, =\, \, (1-s^4)^{1/4} \cdot \tilde{\chi}^{(2n+1)}$ so that
$\, \tilde{\chi}^{(2n+1)}$ has the ``simpler'' divergence
$\propto \, 1/(1-4w) \, \propto \, 1/(1-s)^2$ at
the ferromagnetic critical point.}, modulo a prime, $L_{33}(\tilde{\chi}^{(5)})\, =\, 0$.
Subsequently in Bostan \etal~\cite{High}, the complexity of this ODE was shown to
be reducible to an inhomogeneous equation
\begin{eqnarray}
\label{eq:L24}
L_{24} (\Phi^{(5)}) \,\, = \, \, \,\, E^{(5)},
\end{eqnarray}
where $\, \Phi^{(5)}$ is the linear combination of 5, 3 and 1-particle contributions
\begin{eqnarray}
\label{eq:Phi5}
\, \Phi^{(5)}\, = \,\, \tilde{\chi}^{(5)}\, -\tilde{\chi}^{(3)}/2\, +\, \tilde{\chi}^{(1)}/120.
\end{eqnarray}
The right hand side of (\ref{eq:L24}) which satisfies $\,L_5(E^{(5)}) = \, 0$ is of the form
\begin{eqnarray}
\label{eq:E5}
&&E^{(5)}\, = \, \,
w \cdot [(1-16\, w^2)^3 \cdot P_{4,0} \cdot K^4 \,
+(1-16\, w^2)^2\cdot P_{3,1} \cdot K^3\, E \,
\nonumber \\
&&\quad \quad +(1-16\, w^2)\cdot P_{2,2} \cdot K^2 \cdot E^2
\, + \, P_{1,3}\cdot K\cdot E^3 \,
\nonumber \\
&&\quad \quad +P_{0,4}\cdot E^4]/(1+4w)^6/(1-16\, w^2)^\kappa,
\end{eqnarray}
where $\, K=\, K(4 \, w)$ and $\, E=\, E(4\, w)$ are complete elliptic integrals
and the $\, P_{i,j}=\, P_{i,j}(w)$ are polynomials. The
degree of the polynomials and the denominator power
$\, \kappa$ in (\ref{eq:E5}) depend on the representation of
$\, L_{24}$ in (\ref{eq:L24}). In the case that $\, L_{24}$ is
minimum order 24, $L_{24}$ is of degree 888, while the $\, P_{i,j}$ are
then of degree (at most) 904 with $\, \kappa\, =\,8$. In~\cite{High}
results were only reported for a non-minimum order representation
modulo a prime. Furthermore it was shown that $\, L_{24}$ could be
factored into $\, L_{12}^{(\rm left)}\, L_{12}^{(\rm right)}$ with
$\, L_{12}^{(\rm right)}$ being reducible into several smaller factors,
all but one known in exact arithmetic. The question of whether
$\, L_{12}^{(\rm left)}$ could be factored was left unresolved.
The results reported in~\cite{experim} and~\cite{High} are an impressive
example of what can be obtained by calculation modulo a prime.
However there are limitations. Knowing $\, L_{24}$ and $\, E^{(5)}$ modulo
a prime enables one to deduce only possible local singularities
of $\tilde{\chi}^{(5)}$ and not its global behaviour. For example, one
cannot determine the amplitudes of the singularities in $\tilde{\chi}^{(5)}$
at the ferromagnetic point at $\, s\, =\, 1$. The leading and first correction
term amplitudes were estimated in~\cite{experim} but this was based on
an analysis of the 2000 term exact integer series that had been obtained
from multiple modulo prime series by the Chinese remainder theorem.
In the present paper we determine the minimum order $\, L_{24}$ and $E^{(5)}$
in exact arithmetic. The results can be found on the website~\cite{http}. We have
also constructed the minimum order exact $L_{29}$ defined by
$\, L_{29}(\Phi^{(5)})\, =\,\, 0$ from these results. We have not actually needed
this operator but provide it nevertheless for those who might find it of interest
and in addition we also report in~\cite{http} the exact integer coefficients of
$\tilde{\chi}^{(5)}$ to 8000 terms that we generated directly\footnote[3]{For the
required $\tilde{\chi}^{(3)}$ see \ref{ferro} and references therein.}
from (\ref{eq:L24}). The global information provided by (\ref{eq:L24}) enables
us to prove that $\, L_{12}^{(\rm left)}$ defined by
$\, L_{24}= \, $ $\, L_{12}^{(\rm left)}\, L_{12}^{(\rm right)}$ cannot be factored.
We confirm the 500 digit amplitude of the leading ferromagnetic singularity of
$\tilde{\chi}^{(5)}$ reported by Bailey \etal~\cite{Bailey} and also some
$|s|=\, 1$ circle singularity amplitudes derived by Nickel~\cite{nickel-99,nickel-00}.
An important question left unresolved in~\cite{experim}
was whether the singularity of the ODE for $\tilde{\chi}^{(5)}$
at the zero of the head polynomial at $\, w =\, 1/2$ was in fact a
singularity of $\tilde{\chi}^{(5)}$ on some branch of
the function. Appendix D in~\cite{experim} provided an
example for which not all the zeros of the head
polynomial of the ODE\footnote[1]{We explicitly
exclude zeros associated with apparent singularities.}
satisfied by an integral were Landau
singularities of the integral. However, the integral
for $\tilde{\chi}^{(5)}$ was deemed too complicated
in~\cite{experim} to perform a complete Landau
singularity analysis leaving only the conclusion
that $\, w =\, 1/2$ was very likely not a Landau singularity
of $\tilde{\chi}^{(5)}$. Here we perform some
analytic continuation of $\tilde{\chi}^{(5)}$
beyond the principal disk $ |s|\leq \, 1$ and
onto other branches. Our exploration, while not exhaustive, is complete enough
to show $\tilde{\chi}^{(5)}$ has the singular behaviour
($1-2w)^{7/2}$ with non-vanishing amplitude on an infinite number of branches.
We have not made any progress in identifying the Landau integrand
singularities that give rise to this behaviour.
\vskip .3cm
In section \ref{ode} we describe briefly how we can combine results
reported in~\cite{High} with multiple modulo prime $\, \Phi^{(5)}$ series
of length at most 4000 terms to obtain the exact $\, L_{24}$ and $E^{(5)}$
in (\ref{eq:L24}). This allows us to determine the series expansion of
$\tilde{\chi}^{(5)}$ at the ferromagnetic point by numerical matching of
solutions. The results are reported in \ref{ferro}. Section \ref{proof} is the
outline of the proof that $\, L_{12}^{(\rm left)}$ defined by
$\, L_{24} = \, $ $\, L_{12}^{(\rm left)}\cdot L_{12}^{(\rm right)}$ cannot be
factored. Finally in section \ref{analyt} we describe the analytic continuation
of $\, \Phi^{(5)}$ we have performed to obtain information on the behaviour
of $\,\tilde{\chi}^{(5)}$ at $\, w\, =\, 1/2$.
\section{The ODE for $\, \Phi^{(5)}$ in exact arithmetic}
\label{ode}
\vskip .1cm
We are looking for the unique minimum order $\, L_{24}$
and associated $\, E^{(5)}$ satisfying (\ref{eq:L24})
for a given $\, \Phi^{(5)}$. This is to be done modulo a prime
for enough primes\footnote[5]{ We have found that
about 90 primes $p<\, 2\, ^{15}$ are sufficient.} that the exact integer
$\, L_{24}$ and $\, E^{(5)}$ can be reconstructed
by the Chinese remainder theorem. To generate the required
$\, \Phi^{(5)}$ in (\ref{eq:Phi5}) directly from the defining $\, \tilde{\chi}^{(n)}$
integrals is impractical as the minimum order $\, L_{24}$ is of degree 888
implying $\, 25 \times 889 \, = \, \, 22225$ unknown
coefficients in $\, L_{24}$. Added to this are the five
polynomials of degree 904 in $\, E^{(5)}$, i.e. another
$\, 5 \times 905\, = \, 4525$ terms. Finding all coefficients
is a straightforward linear algebra problem but
requires that we have each $\, \Phi^{(5)}$ modulo prime
series to about 26800 terms\footnote[2]{An alternative
to (\ref{eq:L24}) is $L_{29}( \Phi^{(5)})\, =\, 0$ but this
requires $\, \Phi^{(5)}$ series to
$\, 30 \times 1238\, = \, 37140$ terms. The change
of our ODE problem from homogeneous to inhomogeneous
form is a substantial reduction in complexity.}.
The operator $\, L$ defined by $\, L(\phi) = \, 0$
for a given $\, \phi$ is not unique if one does not require $L$ to
be of minimum order. The advantage of seeking a non-minimal
order $\, L$ is that the number of unknown
coefficients to be found can drop dramatically. For
example, only about 6200 terms are needed
to obtain the non-minimal $\, L_{29}$ of order 51 and
degree 118 used in~\cite{High}. The analogous effect occurs
for the inhomogeneous equation (\ref{eq:L24})
and we find that $\, \Phi^{(5)}$
series of length only about 5300 terms
are needed when $\, L_{24}$ is chosen of order 42 and
degree 103 leading to polynomials of degree 155
in $\, E^{(5)}$ in (\ref{eq:E5}) with now
$\, \kappa \, =\,26$. Note however, as observed
in~\cite{experim}, that the integer coefficients in
non-minimal order operators can be outrageously large
and we expect that Chinese remainder
reconstruction of the non-minimal order $\, L_{24}$ would
be nearly hopeless. Instead, the utility of
a modulo prime non-minimal $\, L_{24}$ lies in its use
as a recursion device to extend directly generated
(short) $ \Phi^{(5)}$ series to series of sufficient length,
i.e. 26800 terms, so that the minimum order $\, L_{24}$ and
associated $\, E^{(5)}$ can be found. Such extension
requires completely negligible computer resources.
A useful variant of the above approach is to use the
non-minimal $\, L_{24}$ found as described above
to generate a long series for a solution $\,S_{24}$
satisfying $\, L_{24}(S_{24})\, =\,\, 0$. This series need only
be about 22300 terms, long enough to enable
reconstruction of the minimum order $\, L_{24}$. Finding
the coefficients in $E^{(5)}$ from $\, L_{24}(\Phi^{(5)})$
is then a separate and simpler problem.
A further reduction in the length of the $\, \Phi^{(5)}$
series to be directly generated can be obtained
if one knows a factorization of $\, L_{24}$ with the right
division operator in exact arithmetic. This
situation can be realized given the four modulo
prime series reported in~\cite{High}. We have
$\, L_{24}\, = \, L_{12}^{(\rm left)}\cdot L_{12}^{(\rm right)}$.
Knowing $\, L_{12}^{(\rm right)}$ in exact arithmetic allows one to
obtain any non-minimal order representation of $\, L_{12}^{(\rm right)}$
modulo any prime\footnote[9]{This is most easily done by using the
minimum order $\, L_{12}^{(\rm right)}$ to generate
a series solution $\,S_{12}$ which
satisfies $\, L_{12}^{(\rm right)}(S_{12}) = \, 0$ but does
not satisfy $\, L_n(S_{12})\, = \,\,0$ for any $ n<12$. The non-minimal
order $\, L_{12}^{(\rm right)}$ is then found from the series
using, for example, the matrix code described in section~3 of~\cite{experim}.}
and then $\Psi\, =$ $\, L_{12}^{(\rm right)}(\Phi^{(5)})$
modulo a prime from the directly generated $\, \Phi^{(5)}$.
If we choose our representation of the known $\, L_{12}^{(\rm right)}$
as order 18 and degree 42, and the unknown $\, L_{12}^{(\rm left)}$ as
order 32 and degree 89, then the polynomials in
$E^{(5)}\, = \,\, L_{12}^{(\rm left)}(\Psi)$ are of degree 199. This
gives $\,33 \times 90\, +\, 5 \times 200\, = \, 3970$
unknown coefficients in $\,E^{(5)}$ and
$\, L_{12}^{(\rm left)}$ to be determined implying that
4000 terms of the directly generated $\,\chi^{(5)}$ series
is more than adequate. Since the series generation of $\,\chi^{(5)}$
described in~\cite{experim} is an
$\, O(N^4\cdot \ln(N))\, $ process, this
represents a 3-fold reduction in computer time
from that needed using an unfactored $\, L_{24}$ or a 6-fold
reduction relative to the unfactored $\, L_{29}$ approach.
We have generated the $\tilde{\chi}^{(5)}$ series
to a minimum $\,\,O(w^{4000})\,$
for 90 primes $\,p\,<\, 2^{15}$ and from these
generated $\, L_{24}$ modulo a prime as described above. To
obtain the exact $\, L_{24}$ is then a problem
of rational reconstruction but it is relatively easy
from just a few terms to guess a
normalization factor\footnote[5]{This becomes the value
of the head polynomial at $w\, =0$ and
is $2^{12}\cdot 3^{13} \cdot 5^7\cdot 7^6\cdot 11^4\cdot 23\cdot 29\cdot 7225564279$
= 4235287273136998077435560752320000000.}
that converts the problem to integer
reconstruction by the Chinese remainder theorem. The
integer coefficients in $\, L_{24}$ are observed
to typically have very large powers of 2 as factors
which one can determine by a process
of trial division by $\, 2^k$. If $\, k$ is chosen too small,
Chinese remainder reconstruction
with a fixed number of primes might fail
because the unknown coefficient is too large
while if $k$ is chosen too large there is failure because
the coefficient is no longer an integer.
With 84 primes we find an intermediate $\, k$
range that yields a consistent integer
reconstruction for every coefficient in $\, L_{24}$. With
90 primes we have a large number
of consistency checks that leave no doubt
that our reconstruction is exact. We have
also confirmed the apparent singularity
constraint equations (A.8) in~\cite{experim}
are satisfied by our reconstructed $\, L_{24}$ in all cases, that is,
19849 satisfied conditions on 22202
non-vanishing coefficients in $L_{24}$.
In view of the ``massive'' calculations required to find $L_{24}$ in exact
arithmetic it is natural to ask for some further mathematical and numerical
checks of the correctness of the operators $\, L_{24}$ and $\, \, L_{12}^{(\rm left)}$.
First of all we have checked directly that $\, \, L_{12}^{(\rm right)}$
does indeed right divide $\, L_{24}$ in exact arithmetic.
Secondly, we have confirmed that the exponents of the operators
$\, L_{24}$ and $\, \, L_{12}^{(\rm left)}$ are {\em rational numbers}
and in agreement with our previous massive numerical calculations~\cite{experim}.
Thirdly, the minimal order operator $\, L_{29}$ corresponds to an integral of an algebraic
integrand, and it is therefore, as mathematicians say a {\em Period}
(``Derived From Geometry''~\cite{bo-bo-ha-ma-we-ze-09}):
$\, L_{29}$ is thus, necessarily, {\em globally nilpotent}. This is a stronger constraint
than being a Fuchsian operator (with integer coefficients) having only rational exponents.
Since $\, L_{24}$ is a factor of $\, L_{29}$ it must also be globally nilpotent,
and likewise the left factor $\, \, L_{12}^{(\rm left)}$ must be globally nilpotent.
We have verified that $\, L_{24}$, $\, \, L_{12}^{(\rm left)}$
and $\, \, L_{12}^{(\rm right)}$\footnote[2]{However, this operator is
automatically globally nilpotent courtesy of its direct sum construction.}
are consistent with globally nilpotent operators. This is a very strong indication
that our exact expressions for $\, L_{24}$ and $\, \, L_{12}^{(\rm left)}$
are correct. To check global nilpotence numerically requires one to
calculate the $p$-curvature and check that it is zero for almost all primes.
In practice one can obviously only do this for the first few primes.
The primes used for $L_{12}$ were those smaller than 30
while primes less than 10 were used for $L_{24}$.
\subsection{Computational details}
As shown in~\cite{experim} the calculation of a series for
$\tilde{\chi}^{(5)}$ is a problem with computational
complexity $O(N^4\ln N)$.
In~\cite{experim} we initially calculated $\tilde{\chi}^{(5)}$ to 10000 terms
which required some 17000 CPU hours on an SGI Altrix cluster with 1.6 GHz
Itanium2 processors. From this single series we could already exactly identify
a simple right divisor of $L_{29},$ and using this factor we were able
to find a solution modulo a second prime using a series of `just' 5600 terms.
Expanding the series to order 5600 took around 1560 CPU hours on the Altrix cluster.
The series modulo these two primes then sufficed to find a larger right divisor
of $L_{29}$ in exact arithmetic, and using this factor we found that only 4800 terms
would be required to find solutions for any subsequent primes.
Shortly after these developments a new system was installed by
the National Computational Infrastructrure (NCI) whose National Facility
provides the national peak computing facility for Australian researchers. This new system is an
SGI XE cluster using quad-core 3.0GHz Intel Harpertown cpus. Our code
runs almost twice as fast on this facility compared to the Altrix cluster.
We then used this system to calculate the series for $\tilde{\chi}^{(5)}$ to order
4800 (this took about 450 CPU hours) for a third prime, which again allowed us to find
an even larger right divisor of $L_{29}$ in exact arithmetic. This larger operator
reduced the required number of terms to 4600
and we then calculated a series for a fourth prime to this order using
some 380 CPU hours. These calculations gave us results for 4 different primes
and allowed us to reconstruct the factor $L_{12}^{(\rm right)}$ in exact arithmetic
as begun in~\cite{High} and completed here in \ref{exactL1tilde}.
It was only after this that we realised that the inhomogenous equation (\ref{eq:L24})
could be used, as detailed above, to obtain simultaneous solutions for
$L_{24}$ and $E^{(5)}$ using as few as 4000 terms. We then calculated series for
$\tilde{\chi}^{(5)}$ to order 4000 for a further 86 primes with each
prime requiring about 215 CPU hours.
\begin{figure}
\begin{center}
\includegraphics[width=12cm]{L24_np.eps}
\end{center}
\caption{\label{fig:L24} Estimates $r_n=\ln (c_n)/\ln (30000)$ for the number of primes
required to reconstruct the coefficients $c_n$.}
\end{figure}
The above timings make it clear that the reconstruction of $L_{24}$ and $\, E^{(5)}$
is a computationally expensive project. The main computational effort is the
direct calculation of the series $\tilde{\chi}^{(5)}$ modulo the required number
of primes. It is therefore of some practical interest to estimate the number
of primes required for the exact reconstruction based only on some partial
reconstruction. We focus here on the coefficients of the head-polynomial of $L_{24}$
and denote by $c_n$ the $n$'th coefficient ($n=0,\ldots,\, 888$)
after stripping it of any factors of 2 as mentioned above. Then $r_n=\ln (c_n)/\ln (30000)$
is a rough measure of the number of primes needed to reconstruct $c_n$. From our previous
reconstruction of $L_{12}^{(\rm right)}$ we noticed that the corresponding
$r_n$ are given roughly by a quadratic function of $n$. Thus one can
estimate the number of primes needed to reconstruct the full $L_{24}$
from a partial reconstruction since $r_n$ for $n$ near 0 or 888 can
be obtained using many fewer primes. In Fig.~\ref{fig:L24} the lower `curve' is
the actual data from the coefficients of the head-polynomial of $L_{24}$,
while the upper solid curves are quadratic fits to the
data based on the part data set, from top to bottom,
$r_n \leq 30,\, 40,\, 50,$ and $60$. The pertinent point being that after
doing the calculation for some 30 primes we estimated that the
reconstruction was likely to succeed with no more than 100 primes and was
therefore achievable in practice.
With $L_{24}$ and $\, E^{(5)}$ known in exact arithmetic it is easy to use
(\ref{eq:L24}) to calculate the exact series for $\Phi^{(5)}$ to high order. Specifically our
reconstruction means that we know the coefficients $a_{i,j}$ of the polynomials
in the operator $L_{24}$ and the coefficients of the polynomials $P_{i,j}$ of (\ref{eq:E5})
exactly. The latter allows us to easily calculate the coefficients $e_n$ of $\, E^{(5)}$
by using the (simple) recursive formulae for the elliptic integrals $E$ and $K$.
From (\ref{eq:L24}) we have explicitly, by equating the coefficients of $x^n$, that
(recall that $L_{24}$ is expressed in terms of the differential operator
$x {\rmd \over \rmd x}$)
\begin{equation} \label{eq:cn}
\sum_{i=0}^M \sum_{j=0}^D a_{i,j}\, (n-j)^i \, c_{n-j} = Ac_n +B = e_n,
\end{equation}
where $A$ and $B$ are integers (depending on $n$).
The coefficients $c_n$ of $\Phi^{(5)}$ can thus be calculated recursively
and from (\ref{eq:Phi5}) we can calculate the coefficients of $\tilde{\chi}^{(5)}$
(with the coefficients of $\tilde{\chi}^{(3)}$ calculated using the ODE from \cite{ze-bo-ha-ma-05c}).
We have calculated the coefficients of $\tilde{\chi}^{(5)}$ up to order 8000 and
they can be found in \cite{http}.
Finally we decided to calculate the minimal order operator $L_{29}$ explicitly;
it is given implicitly by (\ref{eq:L24}). This can be done in a variety of ways.
The most obvious way is to use (\ref{eq:cn}) to extend the series for
$\Phi^{(5)} $ to high enough order ($30\times 1238 = 37140$, since
the minimal order $L_{29}$ has degree 1237) and then use the matrix
code of \cite{experim} to calculate the ODE corresponding to $L_{29}$ modulo
a sufficient number of primes to reconstruct $L_{29}$. However, computationally
it is easier to first calculate modulo a prime the minimal operator $L_5$ annihilating
$E^{(5)}$ and then form the product $L_5\cdot L_{24}$ modulo a prime. The minimal
order operator for $L_5$ has degree $4489$ so $6\times 4490 = 26490$ terms
of $E^{(5)}$ is required. Obviously, since the minimal order $L_{24}$ has
degree 888, the product $L_5\cdot L_{24}$ has degree 5377 meaning that there is
a common factor of degree 4140, which we must discard in order to calculate
the degree 1237 polynomials of $L_{29}$. The minimal order $L_5$ was
calculated (for each prime) using the matrix code of \cite{experim}.
The product $L_5\cdot L_{24}$ was then calculated modulo a prime
using Maple and the common factor can then be divided out modulo a prime.
The bottle-neck in this calculation is the use of the matrix code of \cite{experim}
which has computational complexity $O(N^3)$. It was for this reason that we
chose the ``indirect'' route of going through $L_5$ to get $L_{29}$.
\vskip .5cm
\section{Proof that $\, L_{12}^{(\rm left)}$ does not factorize}
\label{proof}
\vskip .1cm
The factorization of $L_{29}$ defined by $\, L_{29}(\Phi^{(5)}) = \, 0$
described in~\cite{High} relied mostly on the
testing of series $\, S(w)$ generated by $L_{29}(S) = \, 0$
modulo a prime around $\, w=\, 0$. If a particular
series is annihilated by an $L_n$ for order $\, n <\, 29$
then $L_n$ right divides $L_{29}$. Depending
on the singularity exponents, a series
solution\footnote[1]{If the solution contains powers
of $\, \ln(w)$ the $\, S(w)$ here is to be interpreted
as the coefficient of the highest power of $\, \ln(w)$.} $S(w) = \,$
$ w^q \cdot (1\, +\alpha_1 \, w\,+\alpha_2 \, w^2\, + \cdots )$
with fixed $\, q$, might be uniquely determined by $L_{29}(S) =\, 0$
or might contain one or more arbitrary rational coefficients $\alpha_i$.
In the latter case the series may be a generator for a right division
operator only for a particular choice of constants $\alpha_i$ and
the problem then is how these particular values might be found.
If there is only one arbitrary coefficient $\alpha$ in the series $\, S(w)$
an exhaustive search is possible in a modulo prime, $\, p$,
calculation because one need then only test $\, p$ separate series
with $\, \alpha$ an integer satisfying $ 0 \, \le \alpha \, < \, p$.
If there is more than one arbitrary $\, \alpha_i$ in the series $\, S(w)$
such brute force ``guessing'' is no longer practical due to computational
time constraints. It is this that prevented the authors of~\cite{High} from
deciding whether $\, L_{12}^{(\rm left)}$ is factorizable. Let us also note
that we were not able to perform a straight formal calculation factorization
of $\, L_{12}^{(\rm left)}$ using its expression in exact arithmetic and our
attempted calculations failed on a computer with 48 Gb of
memory\footnote[5]{More precisely the Maple 13 command
DFactor(*,onestep) was not able to yield a conclusive answer for such
complicated large order operators. Seeking for a left division of $\, L_{12}^{(\rm left)}$
we had similar inconclusive Maple calculations on $\, {\rm adj}(L_{12}^{(\rm left)})$,
the adjoint operator of $\,L_{12}^{(\rm left)}$.}.
As pointed out in~\cite{High} there is nothing special about the singular point
$w\, =\, 0$. And indeed, a series solution about a singular point $w_s \, \ne \, 0$
might also lead to a right division operator. For example, the unique singular series
\begin{equation}
\label{7over4}
\fl
\qquad x^{-7/4} \cdot (1\, +387\, x/80\, +72103\, x^2/23040\, \
+2054561\, x^3/1597440\, +O(x^4))
\end{equation}
with $\, x=\, w \, -1/4$ is annihilated by an order three operator.
With $\, x=\, w\, -1/2$, the unique singular series
\begin{equation}
\label{7over2}
\fl
\qquad x^{7/2} \cdot (1\, -41\, x/6\, +26557\, x^2/792\, -8692015\, x^3/61776\,\, +\,\, O(x^4))
\end{equation}
yields an order 6 right division operator $\, L_{6}$. By proceeding through a sequence
of such solutions, one can eliminate much if not all of the $p$-fold searching
described in~\cite{High} to achieve the factorization $\, L_ {29} \, = \, L_5 \cdot L_{24}\, =
\, L_5 \cdot L_{12}^{(\rm left)} \cdot L_{12}^{(\rm right)}$
and the additional factorization of $\, L_{12}^{(\rm right)}$.
The points chosen for series expansion need not be restricted to the rational head
polynomial roots as in the equations (\ref{7over4}) and (\ref{7over2}) above.
The factor $\, 1\, +3w\, +4\, w^2$, which appeared already in the head polynomial
of the $\, L_7$ that annihilated $\, {\tilde \chi}^{(3)}$, has ``accidental'' modulo
prime factorizations for roughly half the primes close to $2^{15}$. For example,
with prime $\,p\,=\,\, 32719$, one has $\, 1\,+3w\, +4w^2\, =\, 4\, (w-8973)\, (w-31925)$
modulo $\, p$. We can write $ x=\, w\, -w_p$ with $w_p$ either 8973 or 31925
and obtain solutions about $x =\, 0$ satisfying
$\, L_{24}(S(x)\cdot \ln^2(x)\, +R(x) \cdot \ln(x)\, +Q(x))\, =\,\, 0$
modulo $\, p$, where $\,R$ and $\, Q$ are regular at $\, x =\, 0$ and the series
$\, S(x)\, =\, x\,\, +O(x^2)$ is unique. Testing shows that $\, S(x)$ is annihilated
(modulo $\, p$) by an order 3 operator $L_3(x)$. As we show in general in~\ref{rightdiv},
one can obtain from $L_3(x)$ a right division (modulo $p$) operator $L_3(w)$ and then
from multiple modulo prime calculations, an exact right division $L_3(w)$ by the
Chinese remainder theorem. This reconstructed $L_3(w)$\footnote[2]{It is equivalent
to the product $\, Z_2 \cdot N_1$ which was shown on general grounds
in~\cite{High} to right divide $L_{29}$.} has the factor $\, 1\, +3\, w \, +4\, w^2$ in
its head polynomial in spite of the fact that $w_p$ is clearly not one of the roots
$(-3 \, \pm \rmi\sqrt{7})/8$ modulo $\, p$, of $\, 1\, +3w+4\, w^2$.
Testing at points other than $\, w=\, 0$ also enables one to exclude certain series
solutions as generators of right division operators. For example, the case 3
polynomial\footnote[1]{We follow the notation of Appendix C in~\cite{experim}.}
$\, 1\, -7\, w\, +5\, w^2\, -4\, w^3$ which is a factor of the head polynomial of
$\, L_{24}$ has the modulo prime, $\, p_0\, =\,\, 32749$, factorization
$32745 \cdot (w^2\, +11821\, w\, +10836)(w\, -3635)$. Define $\,x\, =\, w\, -3635$.
Then the singular solution modulo $p_0$ about $x =\, 0\,$ is $\, S(x)\, \ln(x)\, +R(x)$ with
\begin{equation}
\label{eq:lnS}
\fl
\qquad S(x)\, = \, \,\, x^5\, +13877\, x^6\, +9339\, x^7\,
+25021\, x^8\, +21884\, x^9\,\, +\,O(x^{10})
\end{equation}
unique and $\, R(x)$ regular at $x =\, 0$. Testing the series~(\ref{eq:lnS})
shows there is no $L_n(S)\, =\,\, 0$ modulo $\, p_0$ for any $n\, < \,24$.
Thus there is no operator of order less than 24 that has $ x\, = \, w \, - \, 3635$
as a factor (modulo $\, p_0$) of the head polynomial and more generally,
$\, 1\, -7\, w\, +5\, w^2\, -4\, w^3$ as a factor. This also implies that any solution
$\, S$ that is singular at some root of $\, 1\, -7\, w\, +5\, w^2\, -4\, w^3$ and
satisfying $\, L_{12}^{(\rm left)}(S) =\, 0$, cannot also be a solution of an operator
of order less than 12 that right divides $\, L_{12}^{(\rm left)}$. The same conclusion
is reached for the remaining case 3 and both case 4 polynomials for
$\tilde{\chi}^{(5)}$ from Appendix C in~\cite{experim}.
If we only knew $\, L_{24}$ or $\, L_{12}^{(\rm left)}$ modulo prime for a few primes,
the above information about singular solutions associated with case 3 and 4
polynomials would not be particularly useful for finding or excluding factorization.
However, with the exact $\, L_{24}$ available one has global information and can
match series solutions about $ w\, =\,\, 0$ to solutions about other singular points
of $\, L_{24}$. In particular, if one can show that every series solution $\, S(w)$
about $w=\, 0$ satisfying\footnote[4]{Equivalently every solution of $\, L_{24}$
that is not a solution of $\, L_{12}^{(\rm right)}$.} $\, L_{12}^{(\rm left)}(S)\, =\, \, 0$
is singular at some root of the case 3 or 4 polynomials then one has proved that
$\, L_{12}^{(\rm left)}$ does not factorize. Our demonstration that this is the case
uses the singular point $w_s= \, 0.15853 \, \cdots $ which is a root of
$\, 1\, -7\, w\, +5\, w^2\, -4\, w^3$. This is a particularly convenient point as it is the
closest root of the head polynomial of $\, L_{24}$ to both $w=\, 0$ and $ w=\, 1/4$.
We begin the demonstration by studying the two linearly independent solutions
of the form $S_i \, =\, A_i(w) \cdot \ln^3(w)$ plus terms with lower powers of $\, \ln(w)$.
The two series
\begin{eqnarray}
\label{eq:A1}
A_1\, & = &\,\, 9\, w\, +261\, w^3\, +1845\, w^4\, +7046\, w^5\, +42771\, w^6\,
+145980\, w^7\, \nonumber \\
&& +785528\, w^8\, +2536628\, w^9 \, +12800309\, w^{10}\, +38627228\, w^{11}
\,\nonumber \\
&& +187738058\, w^{12}\, +\, \cdots \,\, +\, \alpha_{1,n} \cdot w^n \, + \, \cdots,
\end{eqnarray}
\begin{eqnarray}
\label{eq:A2}
A_2 \, &=& \, \, 27\, w^2 \, +102\, w^{3}\, +270\, w^{4}\, +2164\, w^{5}\, +5532\, w^{6}\,
+43722\, w^{7}\, \nonumber \\
&& +132130\, w^{8}\, +922108\, w^{9}\, +3158590\, w^{10}\, +19690882\, w^{11}\,\nonumber \\
&& +72977164\, w^{12}\, +\, \cdots \, \, +\, \alpha_{2,n} \cdot w^n \, + \, \cdots,
\end{eqnarray}
satisfy $\, L_{24}(A_i)\, =\,\, 0$ {\em but not} $\, L_{12}^{(\rm right)}(A_i)\, =\,\, 0$
and a ratio test shows they have radii of convergence $\, |w| =\, w_s\,=\, 0.15853\cdots$.
Thus both $A_1$ and $A_2$ are singular at $ w= \, w_s$ and cannot
be generators of an $L_n$, $\, n <\, 24$, that right divides $\, L_{24}$. Whether a
linear combination of $A_1$ and $A_2$ leads to a right division
operator is now determined as follows.
Near $ x =\, 0$ where $x \,= \, w_s\, -w$ the $\, A_i$ must be of the form
$A_i\, = \, \, B_i \cdot f(x)\cdot \ln(x)\, +g_i(x)$ where $f(x)$
and the $g_i(x)$ are all regular at $\, x\, =\, 0$. The series
$\, A(w)\, \propto \, B_2 \, A_1(w) \, \, -B_1\, A_2(w)\, $ will be
regular at $ w= \, w_s$ and if the amplitude ratio $B_1/B_2$
is rational, then $A(w)$ might be a
candidate generator for a right division operator. In principle, the amplitudes
$B_i$ could be found by matching series solutions
about $x=\, 0$ to those about $ w=\, 0$ but since
we only need the $B_1/B_2$ ratio, a simpler procedure that utilizes only the
coefficients in (\ref{eq:A1},\ref{eq:A2}) is possible. We note that of the
remaining singularities of $A_i$, the nearest to
$ w=\, 0$ are at $|w|=\,1/4$. This implies
that $\, B_1/B_2\, =\, \alpha_{1,n}/\alpha_{2,n}$, a
ratio of coefficients from (\ref{eq:A1},\ref{eq:A2}), to an accuracy of
order $(4\, w_s)^n \approx \, 0.634^n$. Thus the
problem reduces to searching for a (small) rational
$B_1/B_2$ from a sequence that converges exponentially. This can be done by
expressing $\alpha_{1,n}/\alpha_{2,n}$
as a continued fraction. We observe that, for $\, n$ greater
than some fixed $n_0$, a particular term in the
continued fraction grows exponentially
which is a clear indication that in the limit $n\, \rightarrow \, \infty$
the continued fraction terminates
and is the (small) rational $B_1/B_2\, =\, -637/228$.
Our result for the linear combination series is then
\begin{equation}
\label{Aw}
\fl \qquad A(w)= 2052w +17199 w^2 +124482 w^3 +592650 w^4
+2984956w^5 +O(w^6)
\end{equation}
which is confirmed to have a radius of convergence $|w|\, =\, 1/4$. However,
testing the series (\ref{Aw}) shows it is not annihilated by
any $\, L_n$, $\, \,\, n<\, 24$, and thus we have
eliminated the only possible linear combination
candidate series for a right division operator.
While direct testing is an easy way to exclude $\, A(w)$ in (\ref{Aw}), such direct testing
is impractical\footnote[3]{This was the problem encountered in~\cite{High}.}
as a general method for excluding the many possible series that arise in the remaining
part of our proof. Instead we supplement direct testing by
a method that relies on the global information provided by the exact $\, L_{24}$.
As an illustration of this method, consider again $\, A(w)$. If one matches (\ref{Aw})
to series about $w =\, 1/4$ one finds that $\, A(w)$ contains, as the
leading logarithmic function, $A_s\cdot \ln ^2(y)$ where $2\, y =\, 1\,-4w\, $ and
\begin{equation}
\label{As}
\fl \quad A_s= 1 -21469\, y/640 -1489293\, y^2/81920 +229328363\, y^3/10485760 + O(y^4)
\end{equation}
A ratio test on the coefficients in (\ref{As}) shows $A_s$
has radius of convergence $|y| =\, 1/2\,-2\, w_s$ and
thus is singular at $ w = \, w_s$. This in turn implies that
there are branches of the function $A(w)$
on which it is singular\footnote[2]{These singularities
can be reached, for example,
by following a path along the real axis from $w=\, 0$ to $\, w$ just
less than $\, 1/4$, circling $w=\, 1/4$ any number $\, N$ of times,
and then moving back to $w=\, w_s$ along the real axis. Since $A_s$
multiplies the leading (unique) logarithmic singularity at
$ w=\,1/4$, there cannot be cancellation
of the $w=\, w_s$ singularity of $A_s$ on all branches
distinguished by $\, N$.} at $w=\, w_s$. By forming the linear
combination $\, A(w)\, \propto \, B_2 \, A_1(w) \, -B_1\, A_2(w)$
we only succeeded in forcing an ``accidental'' cancellation
of the singularity at $w =\, w_s$ on the principal
branch of the function. The singularity remains on at least some
other branches and thus $A$ is excluded as a generator of a
right division operator for exactly the same reason as $\, A_1$ and $\, A_2$.
The above argument has excluded, as generators of right division operators,
eight of twelve linearly independent solutions satisfying$\, L_{12}^{(\rm left)}(S) =\, 0$.
These we take to be $C_1\, \propto \, L_{12}^{(\rm right)}(A)$ with $\, A$ from (\ref{Aw})
and $C_2\, \propto \, \, L_{12}^{(\rm right)}(A_2)$ with $A_2$ from (\ref{eq:A2}) plus
the six series with leading $\, \ln(w)$ dependence $\, C_i \cdot \ln^p(w)$,
$p=\, 1,\, 2$ and $3$. Explicitly,
\begin{eqnarray}
\label{C1w}
&& \fl \qquad C_1(w)\, =\,\, w^6\, -444\, w^7/11\,
+275773109\, w^{8}/129360\, +19252320091\, w^{9}/194040 \nonumber \\
&& -964738631897\, w^{10}/388080 -2082457681309\, w^{11}/27720\,\,\nonumber \\
&& +17517580633073581\, w^{12}/17075520\,+\,O(\, w^{13}),
\end{eqnarray}
\begin{eqnarray}
\label{C2w}
&& \fl \qquad C_2(w)\, = \,\,w^6\, +403206\, w^{7}/1661\, -13446782071\, w^{8}/19533360\,
\nonumber \\
&& -2413114741889\, w^{9}/29300040\, -4359267083039 \, w^{10}/1065456\,
\nonumber \\
&& -875856906689449\, w^{11}/4185720\,\nonumber \\
&& +23619065101886078533\, w^{12}/2578403520
\, +\, O(w^{13}).
\end{eqnarray}
Because $\, L_{12}^{(\rm right)}$ does not have the factor $ w\,-w_s\, $ in its
head polynomial, the $\, C(w)$ functions carry the same $w\, =\, w_s$
singularities as the $\,A(w)$ from which they have been generated. Thus
even though the series $C_1(w)$ has radius of convergence $|w|=\, 1/4$,
the analytically continued function $C_1(w)$ is still singular at $w=\, w_s$
on some other branches. The series $C_2(w)$ has radius of convergence
$|w|=\, w_s$ and is already singular at $w=\, w_s$ on the principal branch.
It was shown in~\cite{High} that the remaining four solutions satisfying
$\, L_{12}^{(\rm left)}(S)\, =\,\, 0$ are of the form $C_i(w) \cdot \ln(w)\, +D_i(w)$
and $C_i(w),\,\, i=\, 3,\, 4$, with the $\, C_i$ and $\, D_i$ regular at $\, w=\, 0$.
These two $\,C_i$, together with the two in (\ref{C1w}, \ref{C2w}), are
linearly independent and can all be generated from the coefficient of $\ln(w)$ in
$\, L_{12}^{(\rm right)}(S_{24})$ where $\,S_{24}$ contains four arbitrary
constants and is of the form
$S_{24}\, = \,F(w)\cdot \ln^2(w)\,+G(w)\cdot \ln(w)\,+H(w)$
with $F,\, G$ and $H$ all regular at $w\, =\, 0$. We demand that $\, F$
satisfies\footnote[3]{For the explicit $ F$ in (\ref{F}), it happens that
$\, \rmd F/\rmd\beta_1$ is annihilated by an $\, L_{11}$ and
$\, \rmd F/\rmd\beta_2$ by an $L_9$. This reduction from
$\, L_{12}$ plays no role in our subsequent arguments.}
$\, L_{12}^{(\rm right)}(F) =\, 0$. This guarantees that $\, C(w)\cdot \ln(w)$
is the leading logarithm in $\, L_{12}^{(\rm right)}(S_{24})$ and
simplifies the subsequent analysis. Only $\, F$ and $\,G $
are relevant for determining $\,C$ and a possible choice is
\begin{eqnarray}
\label{F}
&&F \, =\,\,\, \beta_1 \cdot \big(3177 \, w \, -174840\, w^{4}\,
-817828\, w^{5}\, -5829558\, w^{6}\, \nonumber \\
&& \quad
-25983762\, w^{7}\, -142882882\, w^{8}\,
-620769318\, w^{9}\, -3086072424\, w^{10}\,
\nonumber \\
&& \quad -13199839762 \, w^{11}\,
-62214586728\, w^{12}\,
\, +\, O(\, w^{13})\big)\, \nonumber \\
&& \quad +\beta_2 \cdot \big(3177\, w^{3}\, +13803\, w^{4}\, +74932\, w^{5} \,
+287997 \, w^{6}\, \nonumber \\
&& \quad +1265280\, w^{7}\, +4296418\, w^{8}\,
+17162736\, w^{9}\, +48945231 \, w^{10}\,
\nonumber \\
&& \quad +173557768\, w^{11}\, +284486847\, w^{12}\,
+\, O(\, w^{13})\big),
\end{eqnarray}
and
\begin{eqnarray}
\label{G}
&&G\, = \, \, \,
\beta_1 \cdot \big(1604883673\, w^{8}/210\, +2823208099\, w^{9}/105
\nonumber \\
&& \qquad +47115755881\, w^{10}/140
+782148892459\, w^{11}/630+O(\, w^{13})\big)
\nonumber \\
&& \quad -\beta_2 \cdot \big(366106439\, w^{8}/1050+1576821038\, w^{9}/525
\nonumber \\
&& \qquad +9206778909\, w^{10}/350
+1049578781449\, w^{11}/6300+O(\, w^{13})\big)\nonumber \\
&& \quad +\beta_3 \cdot \big(35\, w^{6}\, +1223\, w^{8}\, +1852\, w^{9}\,
+36064\, w^{10}\, \nonumber \\
&& \qquad +96388\, w^{11}\, +O(\, w^{13})\big)\nonumber \\
&& \quad
+\beta_4 \cdot \big(105\, w^{7}\, +304\, w^{8}\, +3536\, w^{9}\,
+10192\, w^{10}\,\nonumber \\
&& \qquad +79089\, w^{11}\,\, +O(\, w^{13})\big),
\end{eqnarray}
where the $\, \beta_i$ are arbitrary constants. We will now show that
no choice of these constants can yield a
$\,C \,= \, C(\beta_1,\, \beta_2,\, \beta_3,\, \beta_4)$, defined by
\begin{eqnarray}
\label{def}
L_{12}^{(\rm right)}(F \cdot \ln^2(w)\,\,+G \cdot \ln(w)\, +H)\,\,
=\,\,\, C \cdot \ln(w)\,\,+D,
\end{eqnarray}
that is a generator for a right division operator of $\, L_{12}^{(\rm left)}$.
The argument is essentially that given above for the exclusion of $C_1$
and $C_2$ in (\ref{C1w}, \ref{C2w}). In fact the demonstration has already
been partially completed since, in terms of
$\, C(\beta_1,\, \beta_2,\, \beta_3,\, \beta_4)$, $\, C_1\, \propto \, C(0,0,3,4)$
and $C_2 \, \propto \, C(0,0,453,1744)$. Since $\, F$ in (\ref{F}) satisfies
$\, L_{12}^{(\rm right)}(F)\, =\,\, 0$ it is not singular at
$w\, = \, w_s=\, 0.15853 \cdots$, neither is $\, L_{12}^{(\rm right)}(F \cdot \ln^2(w))$.
Thus it suffices to investigate $\, G $ and if every $G$ is singular at $\, w =\, w_s$
then so is $\, C$ defined by (\ref{def}) and we have proved $\, L_{12}^{(\rm left)}$
does not factorize.
A ratio test on the series coefficients in (\ref{G}) shows the generic $G$ has
radius of convergence $|w|=\, w_s$ and thus is singular at $ w=\, w_s$.
But by the same analysis that led from the series (\ref{eq:A1}, \ref{eq:A2})
to the linear combination (\ref{Aw}), we can construct three
$G = \, G(\beta_1,\, \beta_2,\, \beta_3,\, \beta_4)$ each of whose radius of
convergence is $\, |w|=\, 1/4$. These are
\begin{eqnarray}
\label{105}
G(105,0,0,-1182781), \quad G(0,525,0,-1443727),\quad G(0,0,3,4).
\end{eqnarray}
To the remaining linearly independent $G =\, G(0,0,0,1)$ one can add any
combination of the three in (\ref{105}) but this will not change its radius
of convergence from $|w|=w_s$ and remove the singularity at $w\, =\, w_s$.
In this sense $ \, G(0,0,0,1)$ is equivalent to $\, G(0,0,453,1744)$
which corresponds to $C_2$ via (\ref{def}) and is excluded as
a generator of any right division operator.
To determine the behaviour of the three $G$ functions in (\ref{105}) in the
vicinity of $y =\, 0$ where $2y\, =\, 1\, -4w$ we match the series\footnote[1]{It follows
from $\, L_{24}(F \cdot \ln^2(w)\, +G \cdot \ln(w)\, +H)\, =\, 0$ by analytic
continuation around the $w=\, 0$ singularity that also
$\, L_{24}(2 \, F \cdot \ln(w)\, +G)= \, 0$.} $ \, 2\, F \cdot \ln(w)\, +\, G$
in $w$ about $w =\, 0$ to solutions $\, S$ satisfying $\, L_{24}(S) =\, 0$
about $y=0$. Since $\, L_{12}^{(\rm right)}(F)=\, 0$ one can show that
$\, F$ (and $F\cdot \ln(w)$) can contain only the first power of $\, \ln(y)$
near $y=\, 0$. Any $\, \ln^2(y)$ or $\, \ln^3(y)$ we find in the matching
$\,S$ can only come from $\, G$ in the combination solution
$\, 2 \, F\cdot \ln(w)\, +G$. Our matching shows the leading logarithmic
dependencies of the three $G's$ in (\ref{105}) are respectively
\begin{eqnarray}
\label{3420025}
&&(-3420025/8192/\pi^2) \, A_s\, \ln^3(y), \quad \quad
(-2473625/16384/\pi^2) \, A_s\, \ln^3(y), \nonumber \\
&& \qquad (-125/16384/\pi^2) \, A_s\, \ln^2(y)
\end{eqnarray}
where $A_s$ is given by (\ref{As}). A linear combination of the first two $G's$
in (\ref{105}) can be constructed to eliminate the leading $A_s \, \ln^3(y)$
shown in (\ref{3420025}) and we find the resultant
$G(494725,-6840050,0,13236968029)$ has $\, A_s\cdot \ln^2(y)$ as the
leading logarithmic singularity. Clearly this remaining singularity can now
be eliminated by forming a linear combination with the last G
in\footnote[5]{The surprise, at least for us, is that all of the necessary
combinations can be formed with rational amplitudes.} (\ref{105}).
In summary, we have generated the three
$\, G(\beta_1,\, \beta_2,\, \beta_3,\, \beta_4)$ combinations
\begin{eqnarray}
\label{494725}
&&G(494725,-6840050,15276842775,33606091729),\quad
\nonumber \\
&&G(0,525,0,-1443727),\quad \quad \quad G(0,0,3,4)
\end{eqnarray}
which are irreducible in the sense that no further superposition can
eliminate the $\, A_s\cdot \ln^3(y)$ and $\, A_s\cdot \ln^2(y)$
singularities of the last two $G's $ while the first $G$ is unique in that
it contains neither $\, A_s\cdot \ln^3(y)$ nor $\, A_s\cdot \ln^2(y)$ .
The last $G$ in (\ref{494725}) we have already identified as being
associated with $C_1$. We associate the middle $\, G$ in (\ref{494725})
with $C_3\propto \, C(0,525,0,-1443727)$ which cannot be a generator
of a right division operator of $\, L_{12}^{(\rm left)}$ for exactly the same
reason as $\, C_1$. The first $\, G$ in (\ref{494725}) we associate with
$C_4\, \propto \, C(494725,-6840050,15276842775,33606091729)$
and test $\, C_4$ directly. We find there is no operator satisfying
$\, L_n(C_4) = \, 0$ with $n <\, 12$ and this completes our proof that
$\, L_{12}^{(\rm left)}$ does not factorize. The explicit new
$C_i$, supplementing those in (\ref{C1w},\ref{C2w}) are
\begin{eqnarray}
&&C_3(w)\, = \,\,w^6+281575923\, w^7/34167793\,
\nonumber \\
&&\quad +48755202697119\, w^8/8371109285
\nonumber \\
&&\quad \, +788146152364265\, w^9/5022665571 \nonumber \\
&&\quad \, -48321460210711729\, w^{10}/33484437140
\nonumber \\
&&\quad \, -1046678480403963299\, w^{11}/4783491020
\nonumber \\
&&\quad \, -4705373665277858926411\, w^{12}/3683288085400\,
+O(\, w^{13}),
\end{eqnarray}
\begin{eqnarray}
&&C_4(w)\, =\,\,w^6\,-124536\, w^{7}/649\,+2840488261\, w^{8}/508816\,
\nonumber \\
&&\quad +39013193251\, w^{9}/254408\,
-24532098411899\, w^{10}/9540300\,
\nonumber \\
&&\quad -6082145734733\, w^{11}/68145\,
\nonumber \\
&&\quad -179169570633725593\, w^{12}/314829900\, +\,O(\, w^{13}).
\end{eqnarray}
\subsection{More on the structure of the differential operator $L_{12}^{\rm (left)}$ }
Once the differential operator $L_{12}^{\rm (left)}$ has been proved to be
irreducible, one may wonder whether this high order differential operator
can nevertheless be built from factors of lower order.
A high order differential operator can be irreducible and still
result from ``operations" involving differential operators of lower order since
it may be a symmetric power of a lower order differential operator
or a symmetric product of two (or more) lower order differential operators.
The symmetric $n$'th power of a differential operator $L_q$ is
the differential operator whose corresponding ODE annihilates a
generic linear combination of the $q$ solutions of $L_q$ to the power $n$.
The symmetric $n$'th power of a differential operator $L_q$ of order $q$
has order ${ (q+n-1)! \over (q-1)!n!}$.
The symmetric product of the differential operators $L_{q_1}$ and $L_{q_2}$
of orders $q_1$ and $q_2$, respectively, is the differential operator whose
ODE annihilates the product of a generic linear combination of the $q_1$ solutions
of $L_{q_1}$ and a generic linear combination of the $q_2$ solutions of $L_{q_2}$.
This symmetric product is of order\footnote[2]{The order $q=q_1 \cdot q_2$ is
for the generic case. In general the order of the symmetric product
is $ q_1 + q_2-1 \le q \le q_1 \cdot q_2$.} $q_1 \cdot q_2$.
We use the notation $[w^p]$ to indicate a series that starts as
$w^p \,(const. + \cdots)$. In \cite{High} it was shown that the formal
solutions of $L_{12}^{\rm (left)}$ at $w=0$ follow this scheme:
There are two sets of four solutions ($k=6,\, 7$)
\begin{eqnarray}
\label{series7}
&& [w^k] \, \ln(w)^3\, + [w^5] \, \ln(w)^2 \,+
[w] \, \ln(w) \, + [w], \, \nonumber \\
&& [w^k] \, \ln(w)^2\, + [w^5] \, \ln(w)\, + [w], \nonumber \\
&& [w^k] \, \ln(w) \, + [w],
\qquad \quad \hbox{and} \quad \qquad [w^k]
\end{eqnarray}
and two sets of two solutions ($k=8,\, 9$)
\begin{eqnarray}
\label{series9}
[w^k] \, \ln(w) \, + [w], \qquad\quad
\hbox{and} \qquad \quad [w^k],
\end{eqnarray}
\begin{eqnarray}
\label{series8}
[w^k] \, \ln(w) \, + [w],
\qquad \quad \hbox{and}\quad \qquad [w^k]
\end{eqnarray}
We denote by $BLn$ a set of solutions such as (\ref{series7}) containing
$n+1$ solutions with a logarithmic solution of maximal degree $n$.
For the scheme above, we thus have two $BL3$ blocks and two $BL1$ blocks,
and in each block there is also a non-logarithmic solution.
We first consider the possibility that $L_{12}^{\rm (left)}$ is a symmetric power of
an operator of lower order. It is straightforward to see that the only possibility
is that $L_{12}^{\rm (left)}$ could be a symmetric eleventh power of a differential
operator of order two with one $BL1$ block.
This possibility is ruled out, since there is no $BL11$ block in
the solutions of $L_{12}^{\rm (left)}$.
Next for the possibility that $L_{12}^{\rm (left)}$ is a symmetric product of
differential operators of lower order (we consider only the cases where
the product has the maximal order). There
are three cases to consider: The symmetric product of differential operators of
orders two and six (configuration denoted $2 \cdot 6$), three and four
(configuration $3 \cdot 4$), or two, two and three (configuration $2 \cdot 2 \cdot 3$).
The symmetric product of two differential operators $L_1$ and $L_2$
containing the blocks $BLn_1$ and $BLn_2$, respectively, should contain in its
solutions the block $BLn$ with $n=n_1+n_2$. Since the differential operator
$L_{12}^{\rm (left)}$ contains two $BL3$ blocks the case $2 \cdot 2 \cdot 3$ is ruled out.
Let us detail the compatibility of the case $3 \cdot 4$ at $w=0$.
With two $BL3$ blocks in $L_{12}^{\rm (left)}$ the only possibility is that
the order three differential must have one $BL2$ block:
\begin{eqnarray}
\label{solBL2}
&& S_1 \, \ln(w)^2 \, + S_{11} \, \ln(w) \, + S_{10}, \nonumber \\
&& S_1 \, \ln(w) \, + S_{20}, \nonumber \\
&& S_1
\end{eqnarray}
and the order four differential operator must have
two $BL1$ blocks:
\begin{eqnarray}
&& T_1 \, \ln(w) \, + T_{10}, \nonumber \\
&& T_1, \\
&& V_1 \, \ln(w) \, + V_{10}, \nonumber \\
&& V_1
\end{eqnarray}
It is a simple calculation to form the product of a combination from the
set $BL2$ with a combination of the solutions from the two sets $BL1$.
One obtains:
\begin{eqnarray}
\fl \quad S_1 \cdot T_1 \, \ln(w)^3 + \left( S_1 \cdot T_{10} + S_{11} \cdot T_1 \right)\, \ln(w)^2
+ \left(S_{10} \cdot T_1 +S_{11} \cdot T_{10} \right) \ln(w)+ S_{10} \cdot T_{10}, \nonumber \\
\label{solST2}
\fl \quad S_1 \cdot T_1\, \ln(w)^2 \, + S_{11} \cdot T_1 \, \ln(w)+ S_{10} \cdot T_1, \\
\label{solST1}
\fl \quad S_1 \cdot T_1\, \ln(w) \, + S_{20} \cdot T_1, \\
\fl \quad S_1 \cdot T_1
\end{eqnarray}
and a similar set of four solutions with $V$'s instead of $T$'s.
These eight solutions correspond to the two $BL3$ occurring for $L_{12}^{\rm (left)}$
at $w=0$. One obtains also
\begin{eqnarray}
\label{solST2p}
&& S_1 \cdot T_1\, \ln(w)^2 \,+ \left( S_1 \cdot T_{10}+T_1 \cdot S_{20} \right) \ln(w)
+ S_{20} \cdot T_{10}, \\
\label{solST1p}
&& S_1 \cdot T_1 \, \ln(w) \,+ S_1 \cdot T_{10}
\end{eqnarray}
and two other solutions where $V$'s replace $T$'s.
Subtracting (\ref{solST2p}) from (\ref{solST2}) and
(\ref{solST1p}) from (\ref{solST1}), one obtains
\begin{eqnarray}
\label{solLast}
&& \left( S_1 \cdot T_{10}+T_1 \cdot S_{20} -T_1 \cdot S_{11} \right)\, \ln(w) \,+
\left( S_{20} \cdot T_{10}-S_{10} \cdot T_1 \right), \\
&& S_1 \cdot T_{10} - T_1 \cdot S_{20}.
\end{eqnarray}
Note that in a set of solutions such as $BL2$ above, the series
$S_{11}$ depends on $S_{20}$ and $S_1$ and can be expressed as
\begin{eqnarray}
S_{11} \,=\, \alpha \, S_1 + 2\, S_{20}
\end{eqnarray}
The coefficient 2 is generic for any order three ODE and $\alpha$ is a constant
that depends on the ODE at hand. Inserting this $S_{11}$ in (\ref{solLast}),
we can arrange to have the non-logarithmic series be the same as the
series in front of the log. This is then one of the $BL1$ blocks occurring in
$L_{12}^{\rm (left)}$. The second $BL1$ block is obtained by considering
(\ref{solLast}) with $V$'s instead of $T$'s.
We have thus shown that under the hypothesis that $L_{12}^{\rm (left)}$ is a symmetric
product of two factors, the scheme of solutions at $w=0$ is compatible with
the configuration $3 \cdot 4$. Similar calculations show that the configuration
$2 \cdot 6$ is also compatible and in this case the order two operator has a $BL1$ block
and the order six operator has two $BL2$ blocks.
Next we must check for each configuration ($3 \cdot 4$ and $2 \cdot 6$)
whether or not the symmetric product is compatible with the scheme of solutions for
$L_{12}^{\rm (left)}$ at those other singularities containing enough logarithmic
solutions. We recall that the scheme at $w=\infty$ is the same as the scheme at $w=0$.
The scheme of solutions of $L_{12}^{\rm (left)}$ at the singularity $w=1/4$ is
one $BL3$ and four $BL1$ blocks. One sees immediately that the configuration $3 \cdot 4$
is ruled out. For any set of solutions that we attach to the order three and order
four differential operators, we end up with either more than one $BL3$ or at
least one $BL2$. The configuration $2 \cdot 6$ is acceptable, since in this case
there can be one $BL1$ block in the order two operator and one $BL2$
plus three $BL0$ blocks in the order six operator.
It remains to be seen whether or not the configuration $2 \cdot 6$ is compatible with the
scheme of solutions at the point $w=-1/4$, which is two $BL2$, one $BL1$ and four
$BL0$ blocks. Our checks show that the configuration $2 \cdot 6$ is ruled out.
In conclusion we have shown that the differential operator $L_{12}^{\rm (left)}$
isn't a symmetric $n'$th power of a lower order operator nor is it
a symmetric product of two (or more) operators of orders $q_1$ and $q_2$ under
the hypothesis that the order of the symmetric product reaches its maximum
value $q_1 \cdot q_2 =12$. It should be noted that our considerations regarding
symmetric powers/products (unlike the question about the factorization of
$L_{12}^{\rm (left)}$) do not require knowledge about $L_{12}^{\rm (left)}$ in
exact arithmetic. The block structure of the solutions can be obtained from
the operator modulo a prime.
\section{Analytic continuation of $\, \Phi^{(5)}$ and its behaviour at $\, w \, = \, 1/2$}
\label{analyt}
\vskip .1cm
Our analytic continuation of $\, \Phi^{(5)}$ is limited to paths that follow the
real $\, w$ axis on the intervals $[0,\, 1/4]$, then $[1/4,\, (3-\sqrt {5})/2]$,
and finally $[(3-\sqrt {5})/2),1/2]$. We allow any number of half-integer turns
about the ferromagnetic point $\, w=\, 1/4$, that is, rotation by any angle
$\, \theta = n\, \pi$ with $\, n$ odd. This is followed by any rotation
$\, \theta = m\, \pi$, $\, m$ odd, around $\, w\, =(3-\sqrt {5})/2$. This point
is one of the $s$-plane circle singularities discussed by Nickel~\cite{nickel-99,nickel-00}.
The point $\, w =\, 1/2$ also maps onto the $|s|=\, 1$ circle but it is not a
singularity of $\tilde{\chi}^{(5)}$ when it is approached on any path that
does not leave the principal disk $|s| \le \, 1$. In terms of our $n,m$ paths
in the $w$-plane, the combinations $n=\, m=\, \pm 1$ are such principal disk
constrained paths. Every other $n,m$ combination is a path that reaches
$\, w=\, 1/2$ on another branch.
Our starting point for the analytic continuation is the $\, \Phi^{(5)}$ series
expansion about $\, w=\, 1/4$ in \ref{ferro}. For $\, 0\, \le\, w <\, 1/4$, $2y=\, 1\,-4w$
is positive real and the half-integer turns about $\, w=\, 1/4$ that bring one to
$ \, w\, >1/4 $ simply requires the replacements
\begin{eqnarray}
\label{yy}
y \, \longrightarrow \, \, -y, \qquad \quad \ln(y/4) \, \longrightarrow
\, \, \, \ln(|y|/4)\, -\rmi n \pi,
\end{eqnarray}
in (\ref{A3}) with $n$ odd to be understood. The new (\ref{A3}) series generated
with the replacements (\ref{yy}) is now to be matched to series in $\, z$ where
$\,z=\, 3\, -\sqrt{5} -2\, w =\, 5/2\, -\sqrt{5} \, -y$. Although direct matching is
possible, considerable improvement in the utility of the $\, y$ series results
by first making an Euler transformation by the replacement $y \rightarrow \, y/(1-y)$.
This has the effect of bringing the $w \, =\, (3\, -\sqrt{5})/2\, \approx \, 0.382$
singular point closer to $\, w=\, 1/4$ while moving $w_s\, \approx \, 0.1585$
further away.
To generate $\, \Phi^{(5)}$ series in $\, z$ we must first analytically continue
the elliptic integrals in (\ref{eq:E5}). The replacements required for $\,w >1/4$,
with the same $\, n$ as in (\ref{yy}), are
\begin{eqnarray}
&&K(4w) \, \longrightarrow \, \, \, u[K(u)\, +\rmi \, n K(u')],
\nonumber \\
\label{eq:Enew}
&&E(4w)\, -K(4w) \, \longrightarrow \, \, \, [E(u)\, -K(u)\, -\rmi \,n E(u')]/u
\end{eqnarray}
where $u=\,\, 1/(4w)$ and $\, u'\, =\, \, \sqrt{1-u^2}$. The new elliptic
integrals (\ref{eq:Enew}) are easily developed as series in $\, z$
from their defining differential equations set up as recursion relations. The remaining
step of finding a particular integral of (\ref{eq:L24}) together with all homogeneous
series solutions is also straightforward. The ODE numerical recursion in $z$ is not
as unstable as the recursion in $\, y$ noted in \ref{ferro}. Here one loses only about
a factor 10 in relative accuracy for each two orders in $\,z$.
The matched series in $z$ is of the form $\, A(z)\cdot \ln(z)\, +B(z)$ where $A$ and $B$
are regular at $z=\, 0$. Since $\tilde{\chi}^{(1)}$ and $\tilde{\chi}^{(3)}$ are not singular
at this point, we can identify the singularity in $ \,\Phi^{(5)} $ with that in $\tilde{\chi}^{(5)}$.
We find the leading singular term is\footnote[2]{This is based on floating point
results to about 300 digit accuracy.}
\begin{equation}
\label{26}
\fl \qquad [(1+10 n^2+5n^4)/16] \cdot [(25/693)(5-\sqrt{5})(2+\sqrt{5})^{11}/(2^{22}\, \pi^2)]
\cdot z^{11} \, \ln(z).
\end{equation}
The $n=\, 1$ singularity, when mapped to $s$-plane variables, agrees with the sum
contribution $\tilde{\chi}^{(5)}_{0,1}\, +\tilde{\chi}^{(5)}_{0,-1}$ from
equation~(14) in~\cite{nickel-99}. The new result in (\ref{26}) is the branch
dependent multiplicity $(1\,+10\, n^2\,+5\,n^4)/16$.
To most clearly identify a possible $\, w\, =\, 1/2$ singularity in the $\, A(z)$ and $\, B(z)$
series we make another Euler transformation with the replacement
$\, z \rightarrow \, z/(1\, +6\,z/5)$. This moves the known singularity at $\, w\, =\, 1/4$
to the new $\, z=\, 5\,(\sqrt{5}-1)/16\, \approx \, 0.386$ and the potential
singularity at $\, w\, =\, 1/2$ to the new $\, z=\, -5\, (16\,-5\, \sqrt{5})/131\, \approx \,-0.184$.
There is another potential singularity at a complex root of the case 4 polynomial
$\,1\,-w\,-3w^2\, +4w^3$. This maps to the new $\, |z|\, \approx \, 0.387$.
Since the singularity of interest is about factor $2.1$ closer than the next nearest,
any singularity at $\, w\, =\, 1/2$ will be observable in an $N$ term Euler transformed
$z$ series with corrections of order $\, (2.1)^{-N}$. This factor $\, (2.1)^{-N}$ is also
the bound we can put on any $\, w\, =\, 1/2$ singularity amplitude if a ratio test
of coefficients does not indicate a singularity at $\, z\, \approx \, -0.184$.
We find the absence of such a singularity for the $A(z)$ series which we have generated
to length $\,N=\, 1100$. Because the $\, A(z)$ series multiplies $\, \ln(z)$, this also
indicates that any possible singularity at $\, w=\, 1/2$ will have an amplitude
independent of the index $\, m$ specifying the logarithmic branch of
the $z=0$ ($w\, =\,(3 -\sqrt{5})/2$) singularity.
We find that for $n =\, \pm 1$, the $\, B(z)$ series is also not singular at
$\, z\, \approx \, -0.184$. For other $n$ values a singularity is clearly indicated
and by a coefficient ratio analysis of different $\, n$ series completely analogous
to what was done for the $\, A_1$ and $\, A_2$ series in (\ref{eq:A1},\ref{eq:A2})
we find that the singularities at $\, w=\, 1/2$ have amplitudes proportional to the
branch dependent multiplicity factor $\,(n^2-1)^2$. A more detailed analysis
involving explicit fitting of the $\, B(z)$ series coefficients yields the singularity
amplitude which we have confirmed by direct matching of the $\, \tilde{\chi}^{(5)}$
series in $z$ to ODE solution series about $\, w=\, 1/2$. The result
for the singular part of $\tilde{\chi}^{(5)}$ at $\, w=\, 1/2$ is
\begin{eqnarray}
\label{finalresult}
&& \tilde{\chi}^{(5)}_{\rm sing}\, =\, \,
\big[(n^2-1)^2 (2\sqrt{6})/(315\, \pi)\big]\cdot\, (1-2\, w)^{7/2}
\nonumber \\
&& \times \, \, [1\, +41\, (1\, -2w )/12\, +26557\, (1-2w)^2/3168\, +\, \cdots ],
\end{eqnarray}
where the amplitude has been verified to our numerical accuracy of
about 250 digits. We have not identified a Landau singularity associated
with (\ref{finalresult}). Finding this Landau singularity in the $\, {\tilde \chi}^{(5)}$
integrand remains as the major unsolved challenge of this paper.
\vskip .1cm
\section{Conclusion}
\label{concl}
\vskip .1cm
We have completed the quest begun in~\cite{experim} for the exact integer
arithmetic ODE satisfied by $ \tilde{\chi}^{(5)}$. While most explicit results are
far too extensive to be published here, a selection can be found on the
website~\cite{http}. These include $\, L_{24}$ and $E^{(5)}$ defined in
(\ref{eq:L24})--(\ref{eq:E5}), the explicit factorization
$\, L_{24} = L_{12}^{(\rm left)}\cdot L_{12}^{(\rm right)}$,
the operator $L_{29}$ and the exact integer high temperature series
for $ \tilde{\chi}^{(5)}$ to 8000 terms generated from (\ref{eq:L24})--(\ref{eq:E5})
and numerical coefficients to 800 digits for the $ \tilde{\chi}^{(5)}$ series at the
ferromagnetic point to supplement (\ref{A4}).
We have used the exact ODE to resolve at least one issue that was left
undecided in~\cite{High}. In particular, we have shown in section~\ref{proof}
that the operator $\, L_{12}^{(\rm left)}$ cannot be factored. The techniques
we have described, both for factorization and proving the converse, are not
all well known and we expect they will find application in other problems.
In addition we have shown that $L_{12}^{\rm (left)}$ isn't a symmetric $n'$th
power of a lower order operator nor is it a symmetric product of two (or more)
operators if the order of the symmetric product reaches its maximum value.
An issue raised in~\cite{experim} has only been partially resolved.
No obvious candidate for a Landau singularity in the integral representation
of $ \tilde{\chi}^{(5)}$ at $\, w=\, 1/2$ could be found and because a Landau
integrand singularity is a necessary condition\footnote{Hwa and
Teplitz (see page 6 in \cite{Hwa}) give \cite{Hadamard}
and \cite{Polk} as the original sources for
the necessary requirement.} \cite{Hwa} for an integral to be singular,
it was conjectured that $ \tilde{\chi}^{(5)}$, necessarily defined by analytic
continuation from its series representation, would not be singular at
$w=1/2$\footnote{The abstract in~\cite{experim} erroneously stated the
absence of a singularity without caveat. In view of the findings of this paper
the conclusion that $ \tilde{\chi}^{(6)}$ is singularity free at $w^2=1/8$ must
now also be discounted.}. We have now, in section~4, shown that
$ \tilde{\chi}^{(5)}$ is singular at $w=1/2$ and so clearly have identified all
the singularities of $ \tilde{\chi}^{(5)}$ by an analysis of the ODE it satisfies.
The $w=1/2$ singularity in $ \tilde{\chi}^{(5)}$ proves the existence of an
associated Landau integrand singularity, but we have made no progress in
identifying this Landau singularity. And so our goal of unifying the Landau
integrand analysis with the ODE approach to the behaviour of the general
$ \tilde{\chi}^{(n)}$ continues to elude us.
\ack
IJ and AJG are supported by the Australian Research Council.
The calculations would not have been possible without a generous grant
from the National Computational Infrastructure (NCI) whose National Facility
provides the national peak computing facility for Australian researchers.
This work has been performed without any support
from the ANR, the ERC, the MAE.
|
\section{Introduction} In \cite{mahan-split} we showed that simply or doubly degenerate surface Kleinian groups without accidental parabolics admit Cannon-Thurston maps,
answering affirmatively a question of Cannon and Thurston (Section 6 of \cite{CT} \cite{CTpub}). In \cite{mahan-elct} we had shown that point pre-images
of the Cannon-Thurston map for simply or doubly degenerate groups without parabolics
correspond to endpoints of leaves of ending laminations whenever a point has more than one pre-image. The main aim of this paper is to apply
the techniques developed in \cite{mahan-split} and the result of \cite{mahan-elct} to extend these results to arbitrary finitely generated
Kleinian groups. This completes the project starting with \cite{mahan-split} and proceeding through \cite{mahan-elct},
\cite{mahan-elct2}, \cite{mahan-red}.
The principal new ingredient is a proof of the existence of Cannon-Thurston maps for degenerate handlebody groups without parabolics.\footnote{An earlier version of
some parts of this paper existed in draft form in an earlier version of \cite{mahan-split}. The division of material
between \cite{mahan-split} and the present paper is in the interests of readability.}
The following is the main new ingredient of this paper. \\
\medskip
\noindent { \bf Theorem \ref{ptpreimagefinal}} {\it Let $G$ be a finitely generated free degenerate Kleinian group without parabolics. Let $i : \Gamma_G \rightarrow {\mathbb H}^3$ be the natural identification
of a Cayley graph of $G$ with the orbit of a point in ${\mathbb H}^3$. Then $i$ extends continuously to a map
$\hat{i}: \hhat{\Gamma_G} \rightarrow {\mathbb D}^3$. Let $\partial i$ denote the restriction of $\hat{i}$ to the boundary $\partial \Gamma$ of $\Gamma$.
Then $\partial i(a) = \partial i(b)$ for $a \neq b \in \partial \Gamma$ if and only if
$a, b$ are either ideal
end-points of a leaf of an ending lamination of $G$, or ideal boundary points of a
complementary ideal polygon. }
\smallskip
A crucial idea in the proof of the existence of the Cannon-Thurston map $\hat i$ in Theorem \ref{ptpreimagefinal} goes back to Miyachi \cite{miyachi-ct} in the case of bounded geometry.
The proof of Theorem \ref{ptpreimagefinal} generalizes with some modifications to arbitrary finitely generated Kleinian groups.
\medskip
\noindent { \bf Theorems \ref{ct} and \ref{ptpreimagefinal-kg}} {\it Let $G$ be a finitely generated Kleinian group.
Let $i : \Gamma_G \rightarrow {\mathbb H}^3$ be the natural identification
of a Cayley graph of $G$ with the orbit of a point in ${\mathbb H}^3$. Then $i$ extends continuously to a map
$\hat{i}: \hhat{\Gamma_G} \rightarrow {\mathbb D}^3$,
where $\hhat{\Gamma_G}$ denotes the (relative) hyperbolic compactification of $\Gamma_G$. Let $\partial i$ denote the restriction of $\hat{i}$ to the boundary $\partial \Gamma_G$ of $\Gamma_G$.
Let $E$ be a degenerate end of $N^h= {\mathbb H}^3/G$ and $\til E$ a lift of $E$ to $\til{N^h}$
and let $M_{gf}$ be an augmented Scott core of $N^h$. Then the ending lamination $\LL_E$ for the end $E$ lifts to a lamination
on $\til{M_{gf}} \cap \til{E}$. Each such lift $\LL$ of the ending lamination of a degenerate end defines a relation $\RR_\LL$ on the (Gromov) Hyperbolic boundary $ \partial \widetilde{M_{gf}}$
(equal to the relative hyperbolic boundary $\partial \Gamma_G$ of $\Gamma_G$),
given by
$a\RR_\LL b$ iff $a, b$ are end-points of a leaf of $\LL$. Let $\{ \RR_i \}_i$ be the entire collection of relations on $ \partial \widetilde{M_{gf}}$ obtained this way. Let $\RR$ be the transitive closure of
the union $\bigcup_i \RR_i$. Then $\partial i(a) = \partial i(b)$ iff $a\RR b$.}
\smallskip
Theorem \ref{ct} gives an affirmative answer to a conjecture of McMullen \cite{ctm-locconn} and Theorem \ref{ptpreimagefinal}
gives an affirmative answer to a conjecture of Otal \cite{otal-thesis}.
For ease of exposition, throughout this paper, we shall often first work out the problem for free groups and then indicate the
generalization to arbitrary finitely generated Kleinian groups.
\smallskip
\noindent {\bf Acknowledgments:} I am grateful to Jean-Pierre Otal for suggesting the problem of finding point pre-images of the Cannon-Thurston
map for handlebodies; and for giving me a copy of his thesis \cite{otal-thesis}, where the structure of Cannon-Thurston maps for handlebody groups
is conjectured. I would also like to thank the referee for several suggestions and corrections.
This work is partly supported by a CEFIPRA Indo-French Research grant 4301-1.
\subsection{Relative Hyperbolicity}
We refer the reader to Farb \cite{farb-relhyp}
for terminology and details on relative
hyperbolicity and electric geometry.
\begin{defn} {\rm
Given a metric space $(X,d_X)$ and a collection $\mathcal{H}$ of subsets, let
$\EE(X,\HH ) = X \bigsqcup_{H \in \HH} (H \times [0,\frac{1}{2}])$ be the
identification space obtained by identifying $(h,0) \in H \times [0,\frac{1}{2}]$
with $h \in X$. Each $\{ h \} \times [0,\frac{1}{2}]$ is declared to be
isometric to the interval $[0, \frac{1}{2} ]$
and $H \times \{ \frac{1}{2} \}$ is equipped with the zero metric. $\EE(X,\HH )$ is given a path pseudo-metric as follows.
Only such paths in $\EE(X,\HH )$ are allowed whose intersection with any $\{ h \} \times (0,\frac{1}{2})$ is either all of $\{ h \} \times (0,\frac{1}{2})$ or is empty.
The distance between two points
in $\EE(X,\HH )$ is the infimum of lengths of such allowable paths.
The resulting pseudo-metric space
$\EE(X,\HH )$ is the {\bf electric space} associated to $X$ and the collection $\HH$. \\
We shall say that $\EE(X,\HH )$ is constructed from $X$ by {\bf electrocuting} the collection $\HH$ and the induced pseudo-metric $d_e$ will
be called the {\bf
electric metric}.\\
If $\EE(X,\HH )$ is (Gromov) hyperbolic, we say that $X$ is {\bf weakly hyperbolic} relative to $\HH$. }
\end{defn}
Note that since $\EE(X,\HH ) = X \bigsqcup_{H \in \HH} (H \times [0,\frac{1}{2}])$, $X$ can be naturally identified with a subspace of $\EE(X,\HH )$.
Paths in $(X,d_X)$ can therefore be regarded as paths in $\EE(X,\HH )$, but are very far from being quasi-isometrically embedded in general.
A collection $\mathcal{H}$ of subsets of $(X,d_X)$ is said to be $D$-separated if $d_X(H_1, H_2) \geq D$
for all $H_1, H_2 \in \HH; H_1 \neq H_2$.
$D$-separatedness is only a technical restriction as the collection $\{ H \times \{ \frac{1}{2} \}: H \in \HH \}$ is $1$-separated in $\EE(X,\HH )$.
\begin{defn} \label{bt} {\rm
Given a collection $\mathcal{H}$
of $C$-quasiconvex, $D$-separated sets in a (Gromov) hyperbolic metric space $(X,d_X)$
\begin{comment}
and a number $\epsilon$
\end{comment}
we
shall say that a geodesic (resp. quasigeodesic) $\gamma$ is a geodesic
(resp. quasigeodesic) {\bf without backtracking}
\begin{comment}
with respect to
$\epsilon -$ neighborhoods
\end{comment}
if $\gamma$ does not return to $H$ after leaving it, for any $H \in \mathcal{H}$. \\
There is a distinguished collection of $1$-separated subsets of $\EE(X,\HH )$ given by $\{ H \times \{ \frac{1}{2} \}: H \in \HH \}$.
An electric quasigeodesic
{\bf without backtracking} in $\EE(X,\HH )$ is an electric quasigeodesic that does not return to $H \times \{ \frac{1}{2} \} $ after leaving it, for any $H \in \mathcal{H}$.
\begin{comment}
A geodesic (resp. quasigeodesic) $\gamma$ is a geodesic
(resp. quasigeodesic) without backtracking if it is a geodesic
(resp. quasigeodesic) without backtracking with respect to
$\epsilon$ neighborhoods for some $\epsilon \geq 0$.
\end{comment}
}
\end{defn}
\smallskip
\noindent {\bf Notation:} For any pseudo metric space $(Z, \rho)$ and $A\subset Z$,
we shall use the notation $N_R(A, \rho) = \{ x \in Z: \rho(x, A) \leq R \}$ as for metric spaces.
\begin{lemma} (Lemma 4.5 and Proposition 4.6 of \cite{farb-relhyp}; Theorem 5.3 of \cite{klarreich}; \cite{bowditch-relhyp})\\
Given $\delta , C$ there exists $\Delta$ such that
if $(X,d_X)$ is a $\delta$-hyperbolic metric space with a collection
$\mathcal{H}$ of $C$-quasiconvex sets.
then,\\
{\rm Electric quasi-geodesics electrically track (Gromov) hyperbolic
geodesics:} For all $P > 0$, there exists $K > 0$ such that if $\beta$ is any electric $P$-quasigeodesic from $x$ to
$y$, and $\gamma$ is a geodesic in $(X,d_X)$ from $x$ to $y$,
then $\beta \subset N_K ( \gamma, d_e )$. Further,\\
1) $\gamma \subset N_K ((N_0 ( \beta, d_e)), d_X)$. \\
2) {\rm Relative Hyperbolicity:}
$X$ is weakly hyperbolic relative to $\HH$. $\EE(X, \HH )$ is $\Delta$-hyperbolic.\\
\label{farb1A}
\end{lemma}
Note that we do not need $D$-separatedness in the hypothesis of Lemma \ref{farb1A}.
Let $(X,d_X)$ be a $\delta$-hyperbolic metric
space, and $\mathcal{H}$ a family of $C$-quasiconvex,
collection of subsets.
Let $\alpha = [a,b]$ be a geodesic in $(X,d_X)$ and $\beta $
an electric
$P$-quasigeodesic without backtracking in $\EE(X, \HH)$
joining $a, b$. Order from the left the collection of maximal subsegments of $\beta$
contained entirely in some $H \times \{ \frac{1}{2} \}: H \in \mathcal{H}$. Let $\{ [p_i, q_i] \times \{ \frac{1}{2} \} ( \subset H_i) \}_i$ be the collection of maximal subsegments.
Replace, as per this order, each path of the form $\{p_i\} \times [0,\frac{1}{2}] \cup [p_i, q_i] \times \{ \frac{1}{2} \} \cup \{q_i\} \times [0,\frac{1}{2}] \subset H_i \times [0,\frac{1}{2}]$
by a geodesic $[p_i,q_i]$ in $X$. The resulting
{\bf connected}
path $\beta_q$ in $X$ is called an {\em electro-ambient representative} of $\beta$ in
$X$.
\begin{lemma} (See Proposition 4.3 of \cite{klarreich}, also see Lemma
3.10 of \cite{mahan-ibdd})
Given $\delta$, $C, P$ there exists $C_3$ such that the following
holds: \\
Let $(X,d_X)$ be a $\delta$-hyperbolic metric space and $\mathcal{H}$ a
family of $C$-quasiconvex
subsets. Let $(X,d_e)$ denote the electric space obtained by
electrocuting elements of $\mathcal{H}$. Then, if $\alpha , \beta_q$
denote respectively a (Gromov) hyperbolic geodesic and an electro-ambient
$P$-quasigeodesic with the same end-points, then $\alpha$ lies in a
(Gromov hyperbolic $d_X-$)
$C_3$ neighborhood of $\beta_q$.
\label{ea-strong}
\end{lemma}
Two paths $\beta , \gamma$ in $(X,d_X)$ with the same endpoints are said to have \emph{similar intersection patterns}
with $\HH$ if there exists $\epsilon >0$, depending only on $(X,\HH)$, such that:
\begin{itemize}
\item {\bf Similar Intersection Patterns 1:} If
precisely one of $\{ \beta , \gamma \}$ meets
some $H \in \mathcal{H}$, then the $d_X$-distance from the entry point
to the
exit point is at most $D$.
\item {\bf Similar Intersection Patterns 2:} If
both $\{ \beta , \gamma \}$ meet some $H \in \mathcal{H}$,
then the distance from the entry point of
$\beta$ to that of $\gamma$ is at most $D$, and similarly for the exit points.
\end{itemize}
\begin{definition} \cite{farb-relhyp} {\rm
Suppose that $X$ is
weakly hyperbolic relative to $\mathcal{H}$.
Suppose that any two electric quasigeodesics without backtracking and with the same endpoints have similar intersection patterns with respect to
the collection $\{ H \times \frac{1}{2} : H \in \mathcal{H} \}$.
Then $(X,\HH)$ is said to satisfy {\bf bounded penetration} and
$X$ is said to be
{\bf strongly hyperbolic} relative to $\mathcal{H}$.}
\end{definition}
The next condition ensures that $(X,\HH)$ is {\bf strongly hyperbolic} relative to $\mathcal{H}$.
\begin{definition} {\rm A collection $\mathcal{H}$ of uniformly
$C$-quasiconvex sets in a $\delta$-hyperbolic metric space $X$
is said to be {\bf mutually D-cobounded} if
for all $H_i, H_j \in \mathcal{H}$, $\pi_i
(H_j)$ has diameter less than $D$, where $\pi_i$ denotes a nearest
point projection of $X$ onto $H_i$. A collection is {\bf mutually
cobounded} if it is mutually D-cobounded for some $D$. }
\end{definition}
\begin{lemma} \cite{mahan-split}
Let $X$be a hyperbolic metric space and $\HH$ a collection of $\epsilon$ neighborhoods of mutually cobounded quasiconvex sets;
then any electro-ambient quasigeodesic is a quasigeodesic in $X$.
\label{ea-genl}
\end{lemma}
\noindent {\bf Partial Electrocution} \\
Let $M$ be a (not necessarily simply connected) convex hyperbolic 3-manifold
with a neighborhood of the cusps excised. Then each boundary component of $M$ is of the form $\sigma \times P$, where $P$ is either
an interval or a circle, and $\sigma$ is a horocycle of some fixed
length $e_0$. In the universal cover $\til M$, if we excise (open) horoballs, we
are left with a manifold whose boundaries are flat horospheres of the
form $\widetilde{\sigma} \times \tilde{P}$. Note that $\tilde{P} = P$ if
$P$ is an interval, and $\mathbb{R}$ if $P$ is a circle (the case for
a $(Z + Z)$-cusp ).
Let $Y$ be a convex simply connected hyperbolic 3-manifold.
Let $\mathcal{B}$ denote a collection of horoballs. Let $X$ denote
$Y$ minus the interior of the horoballs in $\mathcal{B}$. Let
$\mathcal{H}$ denote the collection of boundary horospheres. Then each
$H \in \mathcal{H}$ with the induced metric is isometric to a Euclidean
product $E^{1} \times L$ for an interval $L\subset \mathbb{R}$. Here $E^1$ denotes Euclidean $1$-space.
{\bf Partially electrocute} each
$H$ by giving it the product of the zero metric with the Euclidean metric,
i.e. on $E^{1}$ put the zero metric and on $L$ put the Euclidean
metric. The resulting space is essentially what one would get (in the spirit of \cite{farb-relhyp}) by gluing
to each $H$ the mapping cylinder of the projection of $H$ onto the $L$-factor. Let $d_{pel}$ denote the partially electrocuted pseudometric
on $X$.
The above construction can be done in the base manifold $M$ itself by equipping the boundary component $\sigma \times P$ with
the product of a zero metric in the $\sigma$ direction and the Euclidean metric in the $P$-direction.
\begin{lemma} \cite{mahan-split}
$(X,d_{pel})$ is a (Gromov) hyperbolic metric space.
\label{pel}
\end{lemma}
\subsection{Cannon-Thurston Maps}
Let $(X,{d_X})$ and $(Y,{d_Y})$ be hyperbolic metric spaces.
By
adjoining the Gromov boundaries $\partial{X}$ and $\partial{Y}$
to $X$ and $Y$, one obtains their compactifications
$\widehat{X}$ and $\widehat{Y}$ respectively.
Let $ i :Y \rightarrow X$ denote a proper map.
\begin{definition} Let $X$ and $Y$ be hyperbolic metric spaces and
$i : Y \rightarrow X$ be a proper map.
A {\bf Cannon-Thurston map} $\hat{i}$ from $\widehat{Y}$ to
$\widehat{X}$ is a continuous extension of $i$.
\end{definition}
Lemma 2.1 of \cite{mitra-trees} below gives a necessary and sufficient condition for the existence of Cannon-Thurston maps.
\begin{lemma} \cite{mitra-trees}
A Cannon-Thurston map from $\widehat{Y}$ to $\widehat{X}$
exists iff the following condition is satisfied:\\
Given ${y_0}\in{Y}$, there exists a non-negative function $M(N)$, such that
$M(N)\rightarrow\infty$ as $N\rightarrow\infty$ and for all geodesic segments
$\lambda$ lying outside an $N$-ball
around ${y_0}\in{Y}$ any geodesic segment in $X$ joining
the end-points of $i(\lambda)$ lies outside the $M(N)$-ball around
$i({y_0})\in{X}$.
\label{contlemma}
\end{lemma}
We shall now give a criterion for the existence of Cannon-Thurston maps between relatively hyperbolic spaces.
Let $X$ and $Y$ be strongly hyperbolic relative to the collections $\HH_X$ and $\HH_Y$ respectively. Let $i\colon Y\to X$ be a weakly type-preserving proper embedding, i.e. for $H_Y\in \HH_Y$ there exists $H_X\in \HH_X$ such that $i(H_Y)\subset H_X$ and images of distinct elements of $\HH_Y$ lie in distinct elements of $\HH_X$.
In the Lemma below, we specialize to the case where $X, Y$ are convex simply connected complete hyperbolic manifolds with some disjoint (open) horoballs removed. $\HH_X$ and $\HH_Y$
will denote the resulting horospheres.
\begin{lemma}\label{crit-relhyp} \cite{mj-pal}
A Cannon-Thurston map for a weakly type-preserving proper embedding $i\colon Y \to X$ exists if and only if
there exists a non-negative function $M(N)$ with
$M(N)\rightarrow \infty$ as $N\rightarrow \infty$ such that the
following holds: \\
Suppose $y_0\in Y$, and $\hat \lambda$ in $\widehat{Y} = \EE(Y, \HH_Y)$ is
an electric quasigeodesic segment starting and ending outside horospheres.
If $\lambda^b = \hat \lambda \setminus \bigcup_{K \in \HH_Y} K$
lies outside an $B_N (y_0) \subset Y$,
then for any electric quasigeodesic $\hat \beta$ joining the
end points of $\hat i (\hat \lambda)$ in $\widehat{X}= \EE(X, \HH_X)$,
$\beta ^b = \hat \beta \setminus \bigcup_{H \in \HH_X} H$ lies outside
$B_{M(N)} (i(y_0)) \subset X$.
\end{lemma}
We shall describe this informally as follows: \\
{\it If $\lambda$ lies outside a large ball {\bf modulo horoballs} then so does any geodesic in $X$ joining its endpoints.}
In \cite{mahan-red} we proved the existence of
Cannon-Thurston maps for Kleinian groups corresponding to
pared manifolds whose boundary is incompressible away from cusps.
\begin{definition} A {\bf pared manifold} is a pair $(M,P)$, where $P
\subset \delta M$
is a (possibly empty) 2-dimensional submanifold with boundary such that \\
\begin{enumerate}
\item the fundamental group of each component of $P$ injects into the
fundamental group of $M$
\item the fundamental group of each component of $P$ contains an abelian
subgroup of finite index.
\item any cylinder $C: (S^1 \times I, \delta S^1 \times I) \rightarrow (M,P)$
such that $\pi_1 (C)$ is injective is homotopic {\it rel} boundary into $P$.
\item $P$ contains every component of $\delta M$ which has an abelian subgroup
of finite index.
\end{enumerate}
\end{definition}
A pared manifold $(M,P)$ is said to have {\bf
incompressible boundary}
if each component of $\partial_0 M = \partial M \setminus P$ is
incompressible in $M$.
The following Theorem summarizes the main results of \cite{mahan-split}, \cite{mahan-elct}, \cite{mahan-elct2}, \cite{mahan-red}. It proves the existence of
Cannon-Thurston maps for Kleinian groups corresponding to
pared manifolds whose boundary is incompressible away from cusps. It also describes the structure of these maps in terms of ending laminations.
\begin{theorem} \cite{mahan-red}
Suppose that $N^h \in H(M,P)$ is a hyperbolic structure
on a pared manifold $(M,P)$ with incompressible boundary. Let
$M_{gf}$ denotes a geometrically finite hyperbolic structure adapted
to $(M,P)$. Then the map $i: \widetilde{M_{gf}}
\rightarrow \widetilde{N^h}$ extends continuously to the boundary
$\partial {i}: \partial \widetilde{M_{gf}}
\rightarrow \partial \widetilde{N^h}$.
Let $E$ be a degenerate end of $N^h$ and $\til E$ a lift of $E$ to $\til{N^h}$. Then the ending lamination $\LL_E$ for the end $E$ lifts to a lamination
on $\til{M_{gf}} \cap \til{E}$. Each such lift $\LL$ of the ending lamination of a degenerate end defines a relation $\RR_\LL$ on the (Gromov) hyperbolic boundary $ \partial \widetilde{M_{gf}}$ given by
$a\RR_\LL b$ iff $a, b$ are end-points of a leaf of $\LL$. Let $\{ \RR_i \}_i$ be the entire collection of relations on $ \partial \widetilde{M_{gf}}$ obtained this way. Let $\RR$ be the transitive closure of
the union $\bigcup_i \RR_i$. Then $\partial i(a) = \partial i(b)$ iff $a\RR b$.
\label{ptpre-inde}
\end{theorem}
\section{Split Geometry} \label{min}
\noindent {\bf Split level Surfaces}\\
A {\bf pants decomposition} of a compact surface $S$, possibly with boundary,
is a disjoint collection of
3-holed spheres $P_1, \cdots , P_n$ embedded in $S$ such that $S \setminus \bigcup_i P_i$ is a disjoint collection of non-peripheral
annuli in $S$, no two of which are homotopic.
\begin{comment}
We shall conflate a pants decomposition of $S$ with the collection of
(isotopy classes of) {\it non-peripheral} boundary curves of $P_1, \cdots , P_n$. Thus when we refer to a pair of pants in a pants decomposition
$P_1, \cdots , P_n$ of $S$ we are referring to one of the $P_i$'s, and when we refer to a curve in a pants decomposition
of $S$ we are referring to one of the non-peripheral boundary curves of one of the $P_i$'s.
\end{comment}
Let $N$ be the convex core of a hyperbolic 3-manifold minus an open neighborhood of the cusp(s). Then any end $E$ of $N$ is simply degenerate
\cite{agol-tameness}, \cite{gab-cal}, \cite{canary} and
homeomorphic to $S \times [0, \infty )$, where $S$ is a compact surface, possibly with boundary. A closed geodesic in an end $E$
homeomorphic to $S \times [0, \infty )$ is {\bf unknotted} if it is isotopic in $E$ to a simple closed curve in $S \times \{0 \}$ via the homeomorphism.
A {\bf tube} in an end $E \subset N$ is a regular $R-$neighborhood
$N(\gamma, R)$ of an unknotted geodesic $\gamma$ in $E$.
\begin{comment}
Let $ \theta , \omega $ be positive real numbers.
A neighborhood $N_\epsilon (\gamma )$ of a closed geodesic $\gamma (\subset N)$
is called a $(\theta, \omega)$-thin tube if the length of $\gamma$ is less than $\theta$ and the length of the shortest
geodesic on $\partial N_\epsilon (\gamma )$ is greater than $\omega$.
\end{comment}
Let $\TT$ denote a collection of disjoint, uniformly separated tubes in ends of $N$
such that\\
a) all Margulis tubes in $E$ belong to $\TT$ for all ends $E$ of $N$.\\
b) there exists $\epsilon_0 >0$ such that the injectivity radius $injrad_x(E) > \epsilon_0$ for all $x \in E \setminus \bigcup_{T \in \TT} Int(T)$ and all ends $E$ of $N$.\\
Let $F: N \rightarrow M$ be a bi-Lipschitz homeomorphism and let $M(0)$
be the image of $N \setminus \bigcup_{T \in \TT} Int(T)$ in $M$ under the
bi-Lipschitz homeomorphism $F$. Let $\partial M(0)$ (resp. $\partial M$) denote the boundary of $M(0)$ (resp. $M$). $M$ will be called the
model manifold. The metrics on $M$ and $\til M$ will be denoted by $d_M$.
Let $(Q, \partial Q)$ be the unique hyperbolic pair of pants such that each component
of $\partial Q$ has length one. $Q$ will be called
the {\it standard} pair of pants.
An isometrically embedded copy of $(Q, \partial Q)$ in $(M(0), \partial M(0))$ will be said to be {\it flat}.
\begin{comment}
The product of the unit circle with the unit interval,
$S^1 \times [0,1]$ will be called the {\it standard annulus}.
\end{comment}
\begin{defn} {\rm A {\bf split level surface} associated to a pants decomposition $\{ Q_1, \cdots , Q_n \}$ of $S$ in $M(0) \subset M$
is an embedding $f : \cup_i (Q_i, \partial Q_i) \rightarrow (M(0), \partial M(0))$ such that \\
1) Each $f (Q_i, \partial Q_i)$ is flat \\
2) $f$ extends to an embedding (also denoted $f$) of $S$ into $M$ such that the interior of each annulus component of
$f(S \setminus \bigcup_i Q_i)$ lies entirely in $F(\bigcup_{T \in \TT} Int(T))$. \\
} \end{defn}
Let
$S_{i}^{s}$ denote the union of the collection of flat pairs of pants
in the image of the embedding $S_{i}$.
The class of {\it all} topological embeddings from $S$ to $M$ that agree with a split level surface $f$
associated to a pants decomposition $\{ Q_1, \cdots , Q_n \}$ on
$Q_1 \cup \cdots \cup Q_n$ will be denoted by $[f]$.
We define a partial order $\leq_E$ on the collection of split level surfaces in an end $E$ of $M$ as follows: \\
$f_1 \leq_E f_2$ if there exist $g_i \in [f_i]$, $i=1,2$, such that $g_2(S)$ lies in the unbounded component of $E \setminus g_1(S)$.
A sequence $S_i$ of split level surfaces is said to exit an end $E$ if $i<j$ implies $S_i \leq_E S_j$ and further for all compact subsets $B \subset E$, there exists
$L>0$ such that $S_i \cap B = \emptyset$ for all $i \geq L$.
\begin{definition} A curve $v$ in $S \subset E$ is {\bf $l$-thin} if the core curve of the Margulis tube $T_v (\subset E \subset N)$ has length less than or equal to $l$.
A tube $T\in \TT$ is $l$-thin if its core curve is $l$-thin. A tube $T\in \TT$ is $l$-thick if it is not $l$-thin. \\
A curve $v$ is said to split a pair of split level surfaces $S_i$ and $S_j$ ($i<j$) if $v$ occurs as a boundary curve of
both $S_i$ and $S_{j}$.\\
\begin{comment}
A pair of split level surfaces $S_i$ and $S_j$ ($i<j$) is said to be an {\bf $l$-thin pair} if there exists an $l$-thin curve $v$ such that
$v$ occurs as a vertex
in both $\tau_i$ and $\tau_{j-1}$.\\
A pair of split level surfaces $S_i$ and $S_j$ ($i<j$) is said to be an {\bf $l$-thin pair on a component
domain $D$} if \\
a) $P_i \cap P_j$ contains a pants decomposition of
$S \setminus D$, none of whose curves are $l$-thin. \\
b) There exists a tight geodesic $g_D \in H$ supported on $D$ such that $(g_D, u) \in \tau_k$
for all $i < k < j$, where the multicurve $u$ contains an $l$-thin curve. Conversely if $(g_D, u) \in \tau_k$ for some $k$, then
$i \leq k \leq j$.
(Here $D$ could be $S$ itself.) Further we demand that
the initial and final vertices of $g_D$ consist of curves contained in (the boundary curves of)
$P_i, P_j$ respectively.
A pair of split level surfaces $S_i$ and $S_j$ ($i<j$) is said to be an {\bf $l$-thick pair}
(or an $l$-thick pair on $S$) if no curve $v \in \tau_k$ is $l$-thin for $i < k < j$.\\
\begin{comment}
A pair of split level surfaces $S_i$ and $S_j$ ($i<j$) is said to be an {\bf $l$-thick pair on $(W_1, \cdots W_l)$}, for $(W_1, \cdots W_l)$ a
collection of component domains of $H$, if \\
a) $P_i \cap P_j$ is a pants decomposition of (the non-annular components of)
$S \setminus \bigcup_k W_k$ \\
b) For each $W_k \in \{ W_1, \cdots W_l \}$, there exists $s$ such that $i < s < j$ and $(g_{W_k}, v)\tau_s$ for some $v$ \\
c) Further, if $(g_D, v) \in \tau_k$ for $i < k < j$ and some $D$, and $v$ is not a curve in $P_i \cap P_j$ then $v$ is not $l$-thin.\\
\end{comment}
\end{definition}
The collection of all $l$-thin tubes is denoted as $\TT_l$. The union of all $l$-thick tubes with $M(0)$ is denoted as $M(l)$.
\begin{defn}
A pair of split level surfaces $S_i$ and $S_j$ ($i<j$) is said to be {\bf $k$-separated} if \\
a) for all $x \in S_i^s$,
$d_M(x,S_j^s) \geq k$\\
b)Similarly, for all $x \in S_j^s$, $d_M(x,S_i^s) \geq k$. \end{defn}
\begin{defn} {\rm An $L$-bi-Lipschitz {\bf split surface} in $M(l)$ associated to a pants decomposition $\{ Q_1, \cdots , Q_n \}$ of $S$
and a collection $\{ A_1, \cdots , A_m \}$ of complementary annuli in $S$
is an embedding $f : \cup_i Q_i \bigcup \cup_i A_i \rightarrow M(l)$ such that\\
1) the restriction $f: \cup_i (Q_i, \partial Q_i) \rightarrow (M(0), \partial M(0))$ is a split level surface \\
2) the restriction $f: A_i \rightarrow M(l)$ is an $L$-bi-Lipschitz embedding.\\
3) $f$ extends to an embedding (also denoted $f$) of $S$ into $M$ such that the interior of each annulus component of
$f(S \setminus (\cup_i Q_i \bigcup \cup_i A_i))$ lies entirely in $F(\bigcup_{T \in \TT_l} Int(T))$.}\end{defn}
\noindent {\bf Note:} The difference between a split level surface and a split surface is that the latter may contain
bi-Lipschitz annuli in addition to flat pairs of pants.
\smallskip
We denote split surfaces by $\Sigma_{i}$ to distinguish them from split level surfaces $S_i$.
Let
$\Sigma_{i}^{s}$ denote the union of the collection of flat pairs of pants
and bi-Lipschitz annuli in the image of the split surface (embedding) $\Sigma_{i}$.
\begin{theorem}
Let $N, M, M(0), S, F$ be as above and $E$ an end of $M$. For any $l$ less than the Margulis constant,
let $M(l) = \{ F(x) : {\rm injrad_x} (N) \geq l \}$. Fix a hyperbolic metric on $S$ such that each component of $\partial S$ is
totally geodesic of length one (this is a normalization condition).
There exist $ L_1 \geq 1$, $ \epsilon_1 > 0$, $n \in \natls$,
and a sequence $\Sigma_i$ of $L_1$-bi-Lipschitz, $ \epsilon_1$-separated split surfaces exiting the end $E$ of $M$
such that for all $i$, one of the following occurs: \\
\begin{enumerate}
\item An $l$-thin curve $v$ splits the pair $(\Sigma_i ,\Sigma_{i+1})$, i.e. $v$ splits the associated split level surfaces $(S_i ,S_{i+1})$, which in turn form
an $l$-thin pair.
\item there exists an $L_1$-bi-Lipschitz embedding $$G_i: (S\times [0,1], (\partial S)\times [0,1]) \rightarrow (M, \partial M)$$
such that $\Sigma_i^s = G_i (S\times \{ 0\})$ and $\Sigma_{i+1}^s = G_i (S\times \{ 1\})$
\end{enumerate}
Finally, each $l$-thin curve in $S$ splits at most
$n$ split level surfaces in the sequence $\{ \Sigma_{i} \}$. \label{wsplit}
\end{theorem}
A model manifold $M$ all of whose ends are equipped with a collection of exiting split surfaces satisfying the conclusions of Theorem \ref{wsplit} is said to be equipped
with a {\bf weak split geometry} structure.
Pairs of split surfaces satisfying Alternative (1) of Theorem \ref{wsplit} will be called an $l$-thin pair of split surfaces (or simply a thin
pair if $l$ is understood). Similarly, pairs of split surfaces satisfying Alternative (2) of Theorem \ref{wsplit} will be called an $l$-thick pair
(or simply a thick
pair) of split surfaces.
\begin{comment}
\begin{rmk} \label{wsplitrmk} {\rm
The notion of split surface could be made a bit more general. We might as well require a split surface to be
a (uniformly) bi-Lipschitz embedding of a bounded geometry subsurface of $S$ containing a pants decomposition.
Theorem \ref{wsplit} then summarizes the consequences of the Minsky model that we shall need in this paper.
We have thus constructed the following from the Minsky model:\\
1) A sequence of split surfaces $S^s_i$ exiting the
end(s) of $M$, where $M$ is marked with a homeomorphism to $S \times J$ ($J$ is $\mathbb{R}$ or $[0, \infty )$ according as $M$ is totally or simply degenerate).
$S_i^s \subset S \times \{ i \}$. \\
2) A collection of Margulis tubes $\mathcal{T}$. \\
3) For each complementary annulus of $S^s_i$ with core $\sigma$,
there is a Margulis tube $T$ whose core is freely homotopic to $\sigma$
such that $T$ intersects $S^s_i$ at the boundary. (What this roughly
means is that there is a $T$ that contains the complementary
annulus.) We say that $T$ splits $S^s_i$.\\
4) There exist constants $\epsilon_0 >0, K_0 >1$ such that
for all $i$, either there exists a Margulis tube splitting both $S^s_i$
and $S^s_{i+1}$, or else $S_i (=S^s_i)$ and $S_{i+1} (=S^s_{i+1})$ have injectivity radius bounded
below by $\epsilon_0$ and bound a {\bf thick block} $B_i$, where a thick block is defined to
be a $K_0-$bi-Lipschitz homeomorphic image of $S \times I$. \\
5) $T \cap S^s_i$ is either empty or consists of a pair of
boundary components of $S^s_i$ that are parallel in $S_i$. \\
6) There is a uniform upper bound $n = n(M)$ on the number of surfaces that
$T$ splits. \\
For easy reference later on, a model manifold satisfying conditions (1)-(6) above is said
to have {\bf weak split geometry}.}
\end{rmk}
\end{comment}
\begin{defn} Let $(\Sigma_i^s, \Sigma_{i+1}^s)$ be a thick pair of split surfaces in $ M$.
The closure of the bounded component of
$M \setminus (\Sigma_i^s \cup \Sigma_{i+1}^s)$ between $\Sigma_i^s, \Sigma_{i+1}^s$ will be called a thick block.\end{defn}
Note that a thick block is uniformly bi-Lipschitz to the product $S \times [0,1]$ and that its boundary components are
$\Sigma_i^s, \Sigma_{i+1}^s$.
\begin{defn} Let $(\Sigma_i^s, \Sigma_{i+1}^s)$ be an $l$-thin pair of split surfaces in $M$
and $F(\TT_i)$ be the collection of $l$-thin Margulis tubes that split both $\Sigma_i^s, \Sigma_{i+1}^s$. The closure of the union of the
bounded components of
$M \setminus ((\Sigma_i^s \cup \Sigma_{i+1}^s)\bigcup_{F(T)\in F(\TT_i)} F(T))$ between $\Sigma_i^s, \Sigma_{i+1}^s$ will be called a split block.
Equivalently, the closure of the union of the
bounded components of
$M(l) \setminus (\Sigma_i^s \cup \Sigma_{i+1}^s)$ between $\Sigma_i^s, \Sigma_{i+1}^s$ is a split block. Each connected component a split block is a split component.
\end{defn}
\begin{rmk} {\rm For each lift $\til{K} \subset
\til{M}$ of a split component $K$ of a split block of $M(l) \subset M$,
there are lifts of $l$-thin Margulis tubes that share the boundary of $\til{K}$ in $\til{M}$. Adjoining these lifts to
$\til{K}$ we obtain {\bf extended split components}. Let $\KK^\prime$ denote the collection of extended split components in $\til{M}$.
Denote the collection of split components in $\til{M(l)} \subset \til{M}$ by $\KK$.
Let $\til{M(l)}$ denote the lift of $M(l)$
to $\til M$.
Then the inclusion of $\til{M(l)}$ into $\til{M}$ gives a quasi-isometry between $\EE (\til{M(l)}, \KK)$ and
$\EE (\til{M}, \KK^\prime)$ equipped with the respective electric metrics. This follows from the last assertion of Theorem \ref{wsplit}.
The electric metric on $\EE (\til{M}, \KK^\prime)$ is called the {\bf graph-metric} and
is denoted by $d_G$. The electric space will be denoted as $(\til{M}, d_G)$.
The electric metric on $\EE (\til{M}, \KK \bigcup \TT_l)$ is quasi-isometric to the electric metric on $\EE (\til{M}, \KK^\prime)$, again by the last assertion of Theorem \ref{wsplit}.
The electric space will be denoted as $(\til{M}, d_G^1)$.}
\end{rmk}
\begin{definition} Let $Y \subset \til{N}$ and $X=F(Y)$. $X \subset \til{M}$ is said to
be $\Delta$-graph quasiconvex if for any hyperbolic geodesic $\mu$ joining $a, b \in Y$,
$F(\mu )$ lies inside $N_\Delta (X, d_G) \subset \EE (\til{M}, \KK^\prime)$. \end{definition}
For $X$ a split component in a manifold, define $CH(X) = F(CH(Y))$, where $CH(Y)$
is the convex hull of $Y$ in $\til{N}$, provided the ends of $N$ have no cusps, i.e. $N=N^h$. Else define $CH(X)$ to be the image under
$F$ of $CH(Y)$ minus cusps.
Further, in order to ensure hyperbolicity of the universal cover, we partially electrocute the cusps of $M$ (cf. Theorem \ref{pel}).
Then $\Delta$-graph quasiconvexity of $X$
is equivalent to the condition that $dia_G (CH(X))$ is bounded by $\Delta^{\prime} = \Delta^{\prime}(\Delta )$ as any split component
has diameter one in $(\til{M}, d_G)$.
\begin{comment}
Split components are
quasiconvex (not necessarily uniformly) in the hyperbolic metric,
and uniformly quasiconvex in the graph metric, i.e. we need to
show {\em hyperbolic quasiconvexity} and {\em uniform graph
quasiconvexity} of split components.\\
\end{comment}
A split component $K (\subset E) \subset N$ is incompressible if the map $i_\ast : \pi_1(K) \rightarrow \pi_1(N)$ induced by the inclusion is injective.
Lemma \ref{hypqc-a}, Proposition \ref{gr-qc-free-a} and Proposition \ref{dGhyp-a} below were proved in \cite{mahan-split} for $M$ homotopy equivalent to a surface, where all split
components are automatically incompressible. However the proofs in \cite{mahan-split} require only that the split
components be incompressible in $M$.
\begin{lemma} Let $E$ be a simply degenerate end of a hyperbolic 3-manifold $N$ equipped with
a weak split geometry model $M$.
For $K$ an incompressible split component contained in $E$, let $\tilde{K}$ be a lift to $\til N$. Then there exists $C_0
= C_0(K)$ such that the convex hull of $\tilde K$ minus cusps lies in a
$C_0$-neighborhood of $\tilde K$ in $\til N$. \label{hypqc-a}
\end{lemma}
\begin{prop} If $K$ is an incompressible split component, then $\til{K} $ is
uniformly graph-quasiconvex in $\til{M}$, i.e. there exists $\Delta^\prime$ such that $dia_G(CH(\til{K}))) \leq \Delta^prime$
for all incompressible split components $\til{K} $.
\label{gr-qc-free-a}
\end{prop}
\begin{comment}
\begin{prop} {\bf Uniform Graph Quasiconvexity of Split Components:}
Each (extended) split component $\til K$ is uniformly
graph-quasiconvex in $(\til{M}, d_G)$.
\label{gr-qc}
\end{prop}
Construct
a second auxiliary metric ${\til{N}}_2=(\til{N}, d_{CH})$
by electrocuting
the elements $CH(\til{K})$ of convex hulls of {\it extended split components}. We show that the spaces ${\til{N}}_1 = (\til{N}, d_{G})$ and
${\tilde{N}}_2=(\til{N}, d_{CH})$ are quasi-isometric. In fact we show that the identity map from $\til N$ to itself induces this quasi-isometry after
the two different electrocutions.
\begin{lemma}
The identity map from $\til N$ to itself
induces a quasi-isometry of
${\widetilde{N}}_1$ and ${\widetilde{N}}_2$.
\label{qi12}
\end{lemma}
\end{comment}
\begin{prop} Suppose that all split components of $\til M$ are incompressible. Then $ (\til{M}, d_{G})$ and hence $ (\til{M}, d_{G}^1)$ are Gromov-hyperbolic.
\label{dGhyp-a}
\end{prop}
In fact electro-ambient quasigeodesics in $ (\til{M}, d_{G})$ and $ (\til{M}, d_{G}^1)$ have the following relation.
\begin{lemma}
Let $M$ be the a model of split geometry such that all split components are incompressible. Let $ (\til{M}, d_{G}) (= \EE(\til{M},\KK^\prime ))$ and $ (\til{M}, d_{G}^1)
(= \EE(\til{M},\KK \bigcup \TT_l ))$ be as above.
Given $o \in \til{M}$ and $C_0 > 0 $,
there exists a function $\Theta : \mathbb{N} \rightarrow \mathbb{N}$ satisfying
$\Theta (n) \rightarrow \infty$ as $n \rightarrow \infty$ such that the following holds.\\
For any $a, b \in \til{M}$, let $\beta_{ea}^h $ be an electro-ambient
$C_0-$quasigeodesic without backtracking in $ (\til{M}, d_{G})$ joining $a, b$. Let
$\beta_{ea} = \beta_{ea}^h \setminus \partial \til{M}$ be the part of $\beta_{ea}^h$ lying away from
the (bi-Lipschitz) horospherical boundary of $ \til{M}$. Again, let $\beta_{ea1}^h $ be an electro-ambient
$C_0-$quasigeodesic without backtracking in $ (\til{M}, d_{G}^1)$ joining $a, b$. Let
$\beta_{ea1} = \beta_{ea1}^h \setminus \partial \til{M}$ be the part of $\beta_{ea1}^h$ lying away from
the (bi-Lipschitz) horospherical boundary of $ \til{M}$.
Then $d_M(\beta_{ea}, o) \geq n $ implies that $d_M(\beta_{ea1}, o) \geq \Theta (n)$. Conversely,
$d_M(\beta_{ea1}, o) \geq n $ implies that $d_M(\beta_{ea}, o) \geq \Theta (n)$.
\label{contlemma1}
\end{lemma}
\begin{proof} Let $K^\prime$ be an extended split component in $\KK^\prime$ and $\til{K^\prime}$ denote its universal cover. Let
$\til{K^\prime} = \til{K} \bigcup \til{T^i}$ where $\til{T^i}$ are the universal covers of $l$-thin Margulis tubes abutting the split component
$\til{K}$ in $\til{K^\prime}$. Suppose $K \subset B_i$ the $i$-th split block in an end $E$.
Then $\til{K^\prime}$ is hyperbolic and contained in a $C(i)$-neighborhood of $\til K$. The argument is now a reprise of similar arguments in Section 6 of \cite{mahan-split}.
Hence for all $i$, there exists $C(i)$, such that $\beta_{ea1} \cap \til{B_i}$ lies in a $C(i)$-neighborhood of
$\beta_{ea} \cap \til{B_i}$ in $\til{M}$. Suppose
$d_M(\beta_{ea}, o) \geq n $. Hence, by uniform $k_0$-separatedness of split surfaces,
$d_M(\beta_{ea1} \cap \til{B_i}, o) \geq max(n-C(i), ik_0)$.
Let $D(i) =max_{1\leq j \leq i} C(i)$. Then $d_M(\beta_{ea1}, o) \geq max(n-D(i), ik_0)$ for all $i$. The Lemma follows.
The converse direction is similar.
\end{proof}
We summarize the conclusions of the above propositions below.
\begin{defn} \label{gqc} A model manifold of weak split geometry is said to be of {\bf split geometry} if \\
\begin{enumerate}
\item Each split component $\til{K}$ is
quasiconvex (not necessarily uniformly) in the hyperbolic metric on $\til{N}$. \\
\item Equip $\til{M}$ with the {\it graph-metric} $d_G$ obtained by
electrocuting (extended) split components $\til{K}$. Then the convex hull
$CH( \til{K})$ of any split component $\til{K}$ has uniformly bounded
diameter in the metric $d_G$.
\end{enumerate}
\end{defn}
Hence by Lemma \ref{hypqc-a} and Proposition \ref{gr-qc-free-a} we have the following.
\begin{theorem}
Any degenerate end of a hyperbolic 3-manifold
is bi-Lipschitz homeomorphic to a Minsky model
and hence to a model of split geometry. \label{minsky-split}
\end{theorem}
\section{Free Groups and Finitely Generated Kleinian Groups}
Let $G$ be a free geometrically infinite Kleinian group.
Agol \cite{agol-tameness}, and independently, Gabai and Calegari \cite{gab-cal} have shown that
$N = {\Hyp^3}/G$ is topologically tame, and hence, by work of Canary
\cite{canary}, geometrically tame. Then any manifold $M$ bi-Lipschitz to $N$ is homeomorphic to the
interior of a handlebody with boundary $S$.
\begin{comment}
Further, from the
proof of the Ending Lamination Conjecture for such groups
\cite{minsky-elc3} \cite{bowditch-endinv}, we know that
a neighborhood $E$ of the end of $M$ is bi-Lipschitz homeomorphic to
a Minsky model for a simply degenerate surface group, where the
surface corresponds to $S$. The same conclusion goes through in the
more general situation of an arbitrary finitely generated Kleinian
group.
\end{comment}
More generally, let $G$ be a finitely
generated Kleinian group and $G_f$ be a geometrically finite
Kleinian group, abstractly isomorphic to $G$ via a type-preserving
isomorphism. Let $H$ denote the convex core of ${\Hyp}^3 / G_f$ and let $M$ be bi-Lipschitz homeomorphic to
$N = {\Hyp}^3 / G$. Then there is a natural identification $i: H \rightarrow
M$ of $H$ with the augmented Scott core (i.e. Scott core plus parabolics) of $M$. Let $\til{i}$ indicate
the lift of $i$ to the universal cover. For most of the discussion below, it might be helpful at a first reading to have in mind a free
geometrically infinite Kleinian group without parabolics. We fix this notation for $H, M, S$ throughout this section.
\subsection{The Masur Domain}
\begin{definition} Let $E$ be an end of a hyperbolic manifold such that $E$ is homeomorphic to $ S \times [0, \infty )$ for $S$ a finite area hyperbolic surface.
A map $h : E \rightarrow S \times [0, \infty )$ is said to
be type-preserving, if all and only
the cusps of $E$ are mapped to cusps of $S \times [0, \infty )$.
\label{def-typepres}
\end{definition}
\begin{theorem} \cite{agol-tameness} \cite{gab-cal}
\cite{bowditch-endinv} \cite{minsky-elc3}
Let $G$ be a finitely generated Kleinian group and $M = {\Hyp^3}/G$. Let
$H$ denote an augmented Scott core of $M$. Let $E_1$ be a
geometrically infinite end of
$M \setminus H$. Then $E_1$ is homeomorphic (via a type-preserving
homeomorphism) to a topological product $S \times [0, \infty )$ for a
hyperbolic
surface $S$ of finite area. Further, there exists a neighborhood $E$
of the end corresponding to $E_1$ such that $E$ is bi-Lipschitz
homeomorphic to a Minsky model for $S \times [0, \infty )$ and
hence to a model of split geometry.
\label{free-model}
\end{theorem}
The last part of the last statement follows from Theorem \ref{minsky-split}.
Some ambiguity remains in the statement of Theorem \ref{free-model}
above. This lies in the choice of the ending lamination for $E$ used
to build the Minsky model. Since $i : S \subset E$ is type-preserving,
no parabolic element of $S$ bounds a compressing disk.
Let $\ML(S)$ be the space of measured laminations. Let $\DD(S)$ be the
subset of $\ML(S)$ consisting of weighted unions of disjoint
meridians (i.e. boundaries of compression disks lying on $S$). Let
$cl(\DD(S) )$ denote the closure of $\DD(S)$. Define the {\bf Masur domain}
of $S$ by \\
$\MM \DD (S) = \{ \lambda \in \ML(S) : i(\lambda, \mu ) > 0\}$ for all
$\mu \in cl(\DD (S) )$, provided $S$ has at least two disjoint isotopy classes of compressing disks.
Else, we define \\
$\MM \DD (S) = \{ \lambda \in \ML(S) : i(\lambda, \mu ) > 0$
for any
$\mu $ that is disjoint from a compressing disk $\}$
Now, let $M = H \cup_i E_i$, where $H$ is an augmented Scott core. Let $S = H \cap E$ be one of
the boundary components of $H$. Let $Mod_0(S)$
denote the subgroup of the mapping class group of $S$ generated by
Dehn twists along essential simple closed curves that bound embedded disks in $H$. As is
customary, a prefix $\PP$ will indicate projectivization. $Mod_0(S)$ acts on $\PP\MM\LL(S)$. It was shown
by Otal \cite{otal-thesis} (see also McCarthy and Papadopoulos
\cite{mcpapa}) that under this action, $Mod_0 (S)$ acts properly discontinuously on
$\PP\MM\DD(S) (\subset \PP\MM\LL(S))$ with limit set $P cl(\DD(S)) (\subset \PP\MM\LL(S))$.
The ending lamination is well-defined up to the action of
$Mod_0(S)$ (see \cite{canary}).
\begin{theorem} \cite{canary}
For any finitely generated Kleinian group, the ending lamination
$\lambda$
facing a surface $S$ with a compressing disk lies in the Masur Domain.
\label{elinmd}
\end{theorem}
For our purposes we shall mostly be satisfied with the fact that $E$
is bi-Lipschitz homeomorphic to {\em some} Minsky model, and hence, by Theorem \ref{free-model}.
\subsection{Incompressibility of Split Components}
We would like to show that sufficiently deep within an end $E$, all split
components are incompressible {\em in $M$}. Recall
that splitting tubes correspond to {\em thin Margulis tubes} in the
split geometry model built from the Minsky model.
\begin{comment}
We state and prove
the propositions for free groups without parabolics and tag on remarks
to indicate the generalizations for arbitrary finitely generated
Kleinian groups.
\end{comment}
\begin{prop}
Let $M = H \cup E$. Equip $E$ with a split geometry structure. Then there exists $E_2 \subset E$ such that \\
\begin{enumerate}
\item $E_2$ is homeomorphic to $S \times [0, \infty )$ by a
type-preserving homeomorphism and consists of a union of blocks and tubes from
the split geometry model for $E$.
\item All split components of $E_2$ are incompressible, i.e. if $K$ is
a split component of $E_2$, then the inclusion $i : K \rightarrow M$
induces an injective map $i_\ast : \pi_1 (K) \rightarrow \pi_1 (M)$.
\end{enumerate}
\label{incompressible}
\end{prop}
\noindent {\bf Proof:} Suppose not. Then there exists a sequence of
split components $K_i$ exiting the end $E$ such that
$i_\ast : \pi_1 (K) \rightarrow \pi_1 (M)$ is not injective. Since
$K_i$ are split components, $K_i = S_i \times I$ for some subsurface
$S_i$ of $S$. By the
Loop Theorem (see for instance, Hempel \cite{hempel-book}), there
exist simple closed curves $\sigma_i \subset S_i$ such
that $\sigma_i$ bound embedded topological disks in $H$. Hence
$\sigma_i \in \DD (S)$. Let $T_i$ be a splitting tube bounding
$K_i$. Then $T_i$ has a core curve $\alpha_i$, which in turn corresponds to a simple closed curve on $S$. It
follows that $d_{CC}(\alpha_i, \DD(S)) = 1$, where $d_{CC}$ denotes distance in the curve-complex. Since any such
sequence of curves $\alpha_i$ converges to the ending lamination
$\lambda$ corresponding to the end $E$, it follows that
$\lambda \in cl(\DD(S))$ and hence cannot lie in the Masur
domain. This contradicts Theorem \ref{elinmd}. The proposition follows. $\Box$
\begin{comment}
\begin{rmk}The above proposition and its proof go through verbatim for
arbitrary finitely generated Kleinian groups, including those with parabolics. \label{incompressible-rmk} \end{rmk}
\end{comment}
As an immediate consequence we have the following.
\begin{lemma}
If $K$ is a split component, then $\pi_1(K)( \subset \pi_1(M))$ is
geometrically finite (Schottky, in the absence of parabolics).
\label{hypqc-free}
\end{lemma}
\begin{proof} Follows from Lemma \ref{hypqc-a}. \end{proof}
\begin{prop} If $K$ is a split component, then $\til{K} $ is
uniformly graph-quasiconvex in $\til{M}$.
\label{gr-qc-free}
\end{prop}
\begin{proof} Follows from Proposition \ref{gr-qc-free-a}. \end{proof}
\begin{prop}
$(\til{M}, d_G)$ is a hyperbolic metric space.
\label{graph_metric_hyperbolic}
\end{prop}
\begin{proof} Follows from Proposition \ref{dGhyp-a}. \end{proof}
\smallskip
\subsection{Constructing Quasidisks}\label{qd}
The construction in this subsection may be regarded as a {\it
graph-metrized coarse}
analogue of an unpublished construction due to Miyachi \cite{miyachi-ct} (see also
Souto \cite{souto-ct} ). The main technical difference between Miyachi's construction and ours is that Miyachi
constructs continuous images of disks that actually separate the universal cover $\til M$,
whereas we only construct quasidisks. As a consequence it becomes technically
more difficult for us to prove that quasidisks coarsely separate. This is why we need a special family of paths which we shall call 'admissible
paths' in the next subsection which either intersect or come close to the quasidisks we construct below.
We choose a
collection of essential simple closed curves $\sigma_1 \cdots \sigma_g$ on $S$ bounding
disks $D_1 \cdots D_g$ with neighborhoods $D_i \times ( -\epsilon ,
\epsilon )$ such that each component of $H \setminus
\bigcup_i D_i \times ( -\epsilon ,
\epsilon )$ is either a ball or has incompressible boundary (rel. cusps). Also assume that $\sigma_i$ are geodesics in
the intrinsic metric on $S$. Next, let $E$ be described as a union of
contiguous blocks $B_k$, where each $B_k$ is either a split block, or
a thick block.
Further, let $\partial B_k = S_{k-1} \cup S_k$ with $S_k$ the {\it
upper boundary} and $S_{k-1}$ the {\it lower boundary}. Also let
$S=S_0$ and $\sigma_i = \sigma_{i0}$. Let $\sigma_{ik}$ be the
shortest closed curve in the split metric on $S_k$ (i.e. in the
pseudometric obtained by electrocuting annular intersections of
splitting Margulis tubes with the split level surface $S_k$) homotopic in $E$ to $\sigma_i = \sigma_{i0}$. Let
$\overline{A_i} = D_i \bigcup_k \sigma_{ik} \subset M$ be the union of the disk
$D_i$ and the quasi-annulus $\bigcup_k \sigma_{ik} $. Then any lift
$A_i$ of $\overline{A_i}$ to $\til{M}$ is isometric to $\overline{A_i}$ as $D_i$ is homotopically
trivial and $\sigma_{ik}$ are all freely homotopic to $\sigma_i =
\sigma_{i0} = \partial D_i$.
We want to show that $A_i$ are quasiconvex in $(\til{M}, d_G)$ which
is hyperbolic by Proposition \ref{graph_metric_hyperbolic}.
\medskip
\noindent{\bf qi Rays}\\ Fix a $\sigma$ and the disk $D$ it bounds. Let $A = D \bigcup_k \sigma_{k} \subset M$, where $\sigma_k \subset S_k$.
Lift $\bigcup_k \sigma_{k}$ to the universal cover $\til{\til E}$ of $E$ such that any lift $\til{\sigma_k}$ lies in the universal cover $ \til{\til{S_k}}$ of $S_k$ (We are using this notation to distinguish
from lifts to $\til M$). Let $\lambda_k$ be any such lift $\til{\sigma_k}$.
We then have the following from \cite{mahan-split}.
\begin{lemma} \cite{mahan-split}
There exists $C \geq 0$ such that for $x_{k} \in \lambda_{k}$ there
exists
$x_{k-1} \in \lambda_{k-1}$ with $d_G (x_{k}, x_{k-1}) \leq C$. Similarly
there
exists
$x_{k+1} \in \lambda_{k+1}$ with $d_G (x_{k}, x_{k+1}) \leq C$. Also, for all $k$ there exists $B(k)$ such that for all
$x_{k} \in \lambda_{k}$ there
exists
$x_{k-1} \in \lambda_{k-1}$ with $d_M (x_{k}, x_{k-1}) \leq B(k)$. Hence, for all $n$ and $x \in \lambda_n$,
there exists a $C$-quasigeodesic ray $r$ such that $r(k) \in \lambda_k \subset {\LL}_\lambda$ for all $k$
and $r(n) = x$.
\label{dGqgeod0}
\end{lemma}
Further, by construction of split blocks, $d_G (x_{i}, S_{i-1}) =
1$. Therefore inductively, $d_G (x_{i}, S_{j}) =
|i-j|$. Hence $d_G (x_{i}, x_{j}) \geq
|i-j|$. By construction, $d_G (x_{i}, x_{j}) \leq
C|i-j|$.
Hence, given $p \in \lambda_{i}$ the sequence of points $ x_{n}, n \in \mathbb{N} \cup \{ 0 \}$
with $x_i = p$ gives by Lemma \ref{dGqgeod0} above, a quasigeodesic in
the $d_G$-metric. Such quasigeodesics shall be referred to as {\em $d_G$-quasigeodesic rays}.
After projecting $\til E$ to $\til{M} \setminus \til{H}$ we have the following conclusion.
\begin{lemma}
There exists $C \geq 0$ such that for all $k$ there exists $B_k$ satisfying the following:\\ For all $x_{ik} \in \sigma_{ik}$ there
exists
$x_{i,k-1} \in \sigma_{i,k-1}$ with $d_G (x_{ik}, x_{i,k-1}) \leq C$ and $d (x_{ik}, x_{i,k-1}) \leq B_k$.
\label{dGqgeod}
\end{lemma}
The following Corollary will turn out to be quite useful.
\begin{cor} There exists $C \geq 0$ such that for all $k$ and all $x_{ik} \in \sigma_{ik}$
there exists $q \in \sigma_{i0}$ and a sequence of points $p = x_{ik},
\cdots, x_{i0}=q$ which is a quasigeodesic in $(\til{M}, d_G)$.
\label{dGqgeod1}
\end{cor}
\begin{proof}
By construction of split blocks, $d_G (x_{ik}, S_{k-1}) =
1$.
Hence, given $p \in \sigma_{ik}$ the sequence of points $p = x_{ik},
\cdots, x_{i0}$ gives by Lemma \ref{dGqgeod} above, a quasigeodesic in
the $d_G$-metric lying entirely on $A_i$ joining $p$ to a point $q \in
D_i$.
\end{proof}
We can choose a point $z_i \in D_i$ (quite arbitrarily) and extend any
quasigeodesic constructed as above by adding on a path from $q$ to
$z_i$ lying entirely in $D_i$ and having uniformly bounded length
(This can be done easily as $D_i$ has bounded diameter. )
\begin{prop}
There exists $C_0 \geq 0$ such that each $A_j$ is $C_0$-quasiconvex in
$(\til{M}, d_G)$.
\label{A-qc}
\end{prop}
\noindent {\bf Proof:} By Lemma \ref{dGqgeod} and Corollary \ref{dGqgeod1}
above, it follows that there exist $K \geq 1$ such
that for any two points $p_1, p_2$ in $A_j$ there exist $K$-
quasi-geodesics $\gamma_1, \gamma_2$ to $z_j$. By Proposition
\ref{graph_metric_hyperbolic}
we also have that $(\til{M}, d_G)$ is hyperbolic. Hence any geodesic
$\alpha_i$ ($i = 1,2$) joining $p_i$ to $z_j$ lies in some $K_1$
neighborhood of $\gamma_i$. Further, by hyperbolicity of $(\til{M},
d_G)$, we conclude that a geodesic $\beta$ joining $p_1, p_2$ lies in a
$K_2$-neighborhood of $\alpha_1 \cup \alpha_2$. Hence, finally,
$\beta$ lies in a $(K_1 + K_2)$ neighborhood of $\gamma_1 \cup
\gamma_2 \subset A_j$. Choosing $C_0 = K_1 + K_2$, we are
through. $\Box$
\begin{comment}
\begin{rmk}The above proposition and its proof also go through for
arbitrary finitely generated Kleinian groups without parabolics. For finitely generated Kleinian groups with parabolics we modify the
model by first partially electrocuting the $\integers$-cusps of the space (cf. Section 8 of \cite{mahan-split}) and the proof
of Proposition \ref{A-qc} goes through for any quasidisk $A_k$ corresponding to a compressing disk. \label{A-qc-rmk} \end{rmk}
\end{comment}
The quasidisks constructed above have the following property.
\begin{lemma} Let $M$ be a model manifold of split geometry. Let $Q$ be a Scott core of $M$
and $\{ D_i \}$ a maximal collection of compressing disks in $Q$.
Then
there exists a function $\Theta : \mathbb{N} \rightarrow \mathbb{N}$ satisfying
$\Theta (n) \rightarrow \infty$ as $n \rightarrow \infty$ such that for all $o \in \til{Q}$ the following holds.\\
Let $ D$ be a lift of one of the $D_i$'s to $\til M$ and let $A$ be the quasidisk in $\til M$ constructed
from $D$ as above.
Then $d_M(D, o) \geq n $ implies that $d_M(A, o) \geq \Theta (n)$.
\label{contlemma-qd}
\end{lemma}
\begin{proof} By Lemma \ref{dGqgeod} and Corollary \ref{dGqgeod1}, there exists $b_1, \cdots , b_k, \cdots$
and $z \in D$ such that for all $x_{k} \in \sigma_{k} \subset A$,
$d_M (x_{k}, z) \leq (b_1 + \cdots + b_k) = c_k$(say).
Hence $d_M (x_{k}, 0) \geq (n-c_k)$.
By uniform $\epsilon_0$-separatedness of split surfaces,
$d_M (x_{k}, 0) \geq k\epsilon_0$.
Hence $d_M (x_{k}, 0) \geq max( (n-c_k), k\epsilon_0)$.
Choosing $\Theta (n) $ to be the largest value of $k$ such that $k\epsilon_0 \leq n-c_k$
we are done. \end{proof}
\subsection{Reduction Lemma and Admissible Paths} Before we get into the proof of the existence
of Cannon-Thurston maps, we recall some material from Section 6 of \cite{mahan-split} that will help
streamline the proof.
The next Lemma allows us to apply the criterion for existence of Cannon-Thurston maps
in Lemmas \ref{contlemma}and \ref{crit-relhyp} to electro-ambient quasigeodesics in $\til M$
rather
than hyperbolic geodesics in $\til N$.
Lemma \ref{contlemma2} below is a paraphrasing of what Lemmas 6.8 and 6.9 of \cite{mahan-split} prove.
\begin{lemma} \cite{mahan-split} Let $N$ be the convex core of a complete hyperbolic $3-$manifold $N^h$ minus a neighborhood
of the cusps. Equip each degenerate end with a split geometry structure such that each split component
is incompressible. Let $M$ be the resulting model of split geometry and $F: N\rightarrow M$ be the bi-Lipschitz
homeomorphism between the two. Let $\til{F}$ be a lift of $F$ to the universal covers.
Then for all $C_0 > 0$, and $o \in \til{N}$
there exists a function $\Theta : \mathbb{N} \rightarrow \mathbb{N}$ satisfying
$\Theta (n) \rightarrow \infty$ as $n \rightarrow \infty$ such that the following holds.\\
For any $a, b \in \til{N}\subset \til{N^h}$, let $\lambda^h $ be the hyperbolic geodesic in $\til{N^h}$ joining
them and let $\lambda^h_{thick} = \lambda^h \cap \til{N}$. Similarly let $\beta_{ea}^h$ be an electro-ambient
$C_0-$quasigeodesic without backtracking in $\til M \subset \EE(\til{M}, \KK^\prime )$
joining $\til{F} (a)$, $\til{F} (b)$. Let
$\beta_{ea} = \beta_{ea}^h \setminus \partial \til{M}$ be the part of $\beta_{ea}^h$ lying away from
the (bi-Lipschitz) horospherical boundary of $ \til{M}$.
Then $d_M(\beta_{ea}, \til{F} (o)) \geq n $ implies that $d_{\Hyp^3}(\lambda^h_{thick}, o) \geq \Theta (n)$.
\label{contlemma2}
\end{lemma}
Combining Lemma \ref{contlemma2} with Lemma \ref{contlemma1}, we have the following.
\begin{cor} Let $N$ be the convex core of a complete hyperbolic $3-$manifold $N^h$ minus a neighborhood
of the cusps. Equip each degenerate end with a split geometry structure such that each split component
is incompressible. Let $M$ be the resulting model of split geometry and $F: N\rightarrow M$ be the bi-Lipschitz
homeomorphism between the two. Let $\til{F}$ be a lift of $F$ to the universal covers.
Then for all $C_0 > 0$, and $o \in \til{N}$
there exists a function $\Theta : \mathbb{N} \rightarrow \mathbb{N}$ satisfying
$\Theta (n) \rightarrow \infty$ as $n \rightarrow \infty$ such that the following holds.\\
For any $a, b \in \til{N}\subset \til{N^h}$, let $\lambda^h $ be the hyperbolic geodesic in $\til{N^h}$ joining
them and let $\lambda^h_{thick} = \lambda^h \cap \til{N}$. Similarly let $\beta_{ea}^h$ be an electro-ambient
$C_0-$quasigeodesic without backtracking in $\til M \subset \EE(\til{M}, \KK\bigcup\TT_l )$
joining $\til{F} (a)$, $\til{F} (b)$. Let
$\beta_{ea} = \beta_{ea}^h \setminus \partial \til{M}$ be the part of $\beta_{ea}^h$ lying away from
the (bi-Lipschitz) horospherical boundary of $ \til{M}$.
Then $d_M(\beta_{ea}, \til{F} (o)) \geq n $ implies that $d_{\Hyp^3}(\lambda^h_{thick}, o) \geq \Theta (n)$.
\label{contlemma3}
\end{cor}
Again, combining Lemma \ref{contlemma2} with Lemma \ref{contlemma-qd}, we have the following.
\begin{cor}
Let $N$ be the convex core of a complete hyperbolic $3-$manifold $N^h$ minus a neighborhood
of the cusps. Equip each degenerate end with a split geometry structure such that each split component
is incompressible. Let $M$ be the resulting model of split geometry and $F: N\rightarrow M$ be the bi-Lipschitz
homeomorphism between the two. Let $\til{F}$ be a lift of $F$ to the universal covers. Let $Q$ be a Scott core of $N$
and $\{ D_i \}$ a maximal collection of compressing disks in $Q$.
Then
there exists a function $\Theta : \mathbb{N} \rightarrow \mathbb{N}$ satisfying
$\Theta (n) \rightarrow \infty$ as $n \rightarrow \infty$ such that for all $o \in \til{Q}$ the following holds.\\
Let $ D$ be a lift of one of the $D_i$'s to $\til M$ and let $A$ be the quasidisk in $\til M$ constructed
from $D$ as above. For any $a,b \in A$, let $[a,b]_h$ be the hyperbolic geodesic in $\til{ N^h}$ joining
${\til{F}}^{-1}(a), {\til{F}}^{-1}(b)$ and let $[a,b] = [a,b]_h\cap \til{N}$.
Then $d_M(D, \til{F}(o)) \geq n $ implies that $d_{\Hyp^3}([a,b], o) \geq \Theta (n)$.
\label{contlemma-qd1}
\end{cor}
\begin{proof} By Lemma \ref{contlemma-qd} there exists $z \in D$, a function
$\Theta_0 : \mathbb{N} \rightarrow \mathbb{N}$ satisfying
$\Theta_0 (n) \rightarrow \infty$ as $n \rightarrow \infty$ and electro-ambient quasigeodesics
$\beta_a, \beta_b$ in $(\til{M},d_G)$ such that $d_M(D, \til{F}(o)) \geq n $ implies that
$d_M(\beta_a \cup \beta_b, \til{F}(o)) \geq \Theta_0 (n)$.
Let $[a,z]_h, [b,z]_h$ be the hyperbolic geodesic in $\til N$ joining
${\til{F}}^{-1}(a), {\til{F}}^{-1}(b)$ respectively to ${\til{F}}^{-1}(z)$
and let $[a,z] = [a,z]_h\cap \til{N}$, $[b,z] = [b,z]_h\cap \til{N}$.
Then by Lemma \ref{contlemma2} there exists a function
$\Theta_1 : \mathbb{N} \rightarrow \mathbb{N}$ satisfying
$\Theta_1 (n) \rightarrow \infty$ as $n \rightarrow \infty$ such that
$d_M(D, \til{F}(o)) \geq n $ implies that $d_{\Hyp^3}([a,z]\cup [b,z], o) \geq \Theta_1 (n)$.
Let $\delta >0$ be such that all geodesic triangles in ${\mathbb{H}}^3$ are $\delta -$thin.
Taking $\Theta (n)=\Theta_1 (n)-\delta$, it follows that
$d_M(D, \til{F}(o)) \geq n $ implies that $d_{\Hyp^3}([a,b], o) \geq \Theta (n)$.
\end{proof}
\noindent {\bf Admissible Quasigeodesics}\\
We shall need a collection of paths consisting of horizontal and vertical segments
approximating electro-ambient quasigeodesics. We shall call these admissible quasigeodesics.
Let $M$ be a model manifold each of whose ends is equipped with a split geometry structure such that all split
components are incompressible. Recall that
each thick block and each split
block in $M$ is homeomorphic to a product $\Sigma_i^s \times I$. We fix such a product structure
for each block.
Let $t_i = sup\{ length(\{x\} \times I): x \in \Sigma_i^s\}$ be the thickness of the $i-$th block.
Recall that $F: N \rightarrow M$ is a bi-Lipschitz homeomorphism from a hyperbolic
manifold $N$ (minus cusps) to $M$ and let $\til F$ denote its lift to the universal cover.
An {\bf elementary admissible path} in $\til M$ is one of the following:
\begin{enumerate}
\item A `horizontal' geodesic in the intrinsic path metric on some lift $\til{\Sigma_i^s}$ of a split surface to $\til M$.
\item A `vertical' path of the form $x \times I$ (with respect to the fixed product
structure above) in the lift to $\til M$ of
either a thick block or a split block.
\end{enumerate}
Let $B_i$ be a thick block of $M$ and let $S_i^s, S_{i+1}^s$ be its horizontal boundary components.
Recall that a product structure $B_i = S \times I$ has been fixed.
Let $\mu = [a,b]$ be a geodesic in $\til{B_i} \subset \til{M}$ such that its end-points
$a, b$ lie on the horizontal boundary components. Let $P(\mu)$ denote the projection of $\mu$
onto the horizontal boundary component ( $S_i^s$ or $ S_{i+1}^s$) containing $a$. If $b$ belongs to
the same horizontal boundary component as $a$, define $\mu_{adm} = P(\mu)$. Else define
$\mu_{adm} = P(\mu)\cup \{b\}\times I$, where $\{b\}\times I$ is the elementary vertical path
through $b$. $\mu_{adm}$ will be called the {\bf admissible quasigeodesic} corresponding
to $\mu$ in the thick block $\til{B_i}$.
Let $\BB$ denote the collection of thick blocks.
\begin{comment}
\item A path of the form $F(\gamma )$ for $\gamma$ a hyperbolic geodesic in $\til N$ lying in the complement
of split components and thick blocks.
Note that paths $\gamma$ in type (3) above are either hyperbolic geodesics inside the lift of an
$l-$thin Margulis tube or in the lift of the Scott core of $N$.
A {\bf $C-$admissible path} in $\til M$ is a path $\beta$ that is
\begin{enumerate}
\item
for any split component $\til K$, $\beta \cap \til{K}$ has at most one vertical
path of type (2) above
\item for any split component $\til K$, and any connected horizontal boundary
component $\til{\Sigma_0^s}$ of $\til K$,
$\beta \cap \til{\Sigma_0^s}$ has at most one `horizontal' geodesic of type (1) above.
\item for any thick block $B$, each component of $\beta \cap \til{B}$ is a $C-$quasigeodesic in $\til M$.
\end{enumerate}
\end{comment}
\begin{defn}
An {\bf admissible quasigeodesic} $\beta_{adm}$ in $\til M$ corresponding to an electro-ambient
$C_0-$quasigeodesic $\beta_{ea}$ without backtracking in $\til M \subset \EE(\til{M}, \KK\bigcup\TT_l )$ is a path such that
$\beta_{adm} \cap (\til{M} \setminus \bigcup_{K \in \KK} \til{K}
\cup \bigcup_{B\in \BB} \til{B} ) = \beta_{ea} \cap (\til{M} \setminus
\bigcup_{K \in \KK} \til{K}\cup \bigcup_{B\in \BB} \til{B} )$.
Further,
for each $ \til{K}$, $\beta_{adm} \cap \til{K}$ is a union of elementary admissible paths with
disjoint interiors such that
\begin{enumerate}
\item $\beta_{adm} \cap \til{K}$ has at most one vertical
path of type (2) above.
\item for any connected horizontal boundary
component $\til{\Sigma_0^s}$ of $\til K$,
$\beta \cap \til{\Sigma_0^s}$ has at most one `horizontal' geodesic of type (1) above.
\end{enumerate}
Finally for each $B\in \BB$, $\beta_{adm} \cap \til{B}$ is the admissible quasigeodesic
corresponding to $\beta_{ea} \cap \til{B}$ in $\til{B}$.
\end{defn}
The next Lemma allows us to apply the criterion for existence of Cannon-Thurston maps
in Lemma \ref{contlemma} to admissible quasigeodesics in $\til M$
rather
than electro-ambient quasigeodesics in $\til M$. The proof of
Lemma \ref{contlemma4} is exactly like Lemma 6.5 of \cite{mahan-split} and we omit it here
(see also the proof of Lemma \ref{contlemma1} above).
\begin{lemma} \cite{mahan-split} Let $N$ be the convex core of a complete hyperbolic $3-$manifold $N^h$ minus a neighborhood
of the cusps. Equip each degenerate end with a split geometry structure such that each split component
is incompressible. Let $M$ be the resulting model of split geometry.
Then for all $C_0 > 0$, and $o \in \til{M}$
there exists a function $\Theta : \mathbb{N} \rightarrow \mathbb{N}$ satisfying
$\Theta (n) \rightarrow \infty$ as $n \rightarrow \infty$ such that the following holds.\\
For any $a, b \in \til{M}$, let $\beta_{ea}^h $ be an electro-ambient
$C_0-$quasigeodesic without backtracking in $\til M$ joining $a, b$. Let
$\beta_{ea} = \beta_{ea}^h \setminus \partial \til{M}$ be the part of $\beta_{ea}^h$ lying away from
the (bi-Lipschitz) horospherical boundary of $ \til{M}$. Again, let $\beta_{adm}^h $ be the admissible
quasigeodesic corresponding to $\beta_{ea}^h $ and let
$\beta_{adm} = \beta_{adm}^h \setminus \partial \til{M}$ be the part of $\beta_{adm}^h$ lying away from
the (bi-Lipschitz) horospherical boundary of $ \til{M}$.
Then $d_M(\beta_{ea}, o) \geq n $ implies that $d_M(\beta_{adm}, o) \geq \Theta (n)$.
Conversely, $d_M(\beta_{adm}, o) \geq n$ implies that $d_M(\beta_{ea}, o) \geq \Theta (n)$.
\label{contlemma4}
\end{lemma}
\subsection{Cannon-Thurston Maps for Free Groups}
We identify $H$ with its bi-Lipschitz image in $M$ under the bi-Lipschitz homeomorphism $F: N\rightarrow M$ from the hyperbolic manifold
$N$ to the model manifold $M$.
We now want to show that if $\lambda = [a,b]$ is a geodesic in the
intrinsic metric on $\til{H}$ joining $a, b \in \til{H}$, and lying
outside a large ball about a fixed reference point $p \in \til{H}
\subset \til{M}$, then the (bi-Lipschitz) hyperbolic geodesic
$\lambda_h$ joining $a, b \in \til{M}$ also lies outside a large ball
about $p$ in $\til{M}$. This would guarantee the existence of a
Cannon-Thurston Map by Lemma \ref{contlemma}.
To fix notation,
let $F$ be a free Kleinian group without parabolics. Let $M = {{\mathbb{H}}^3}/G$
and $H$ a compact (Scott) core of $M$. $\til{H}$ with its intrinsic
metric is quasi-isometric to the Cayley graph $\Gamma_F$ and so its
intrinsic boundary may be identified with the Cantor set $\partial F$
thought of as the Gromov boundary of $\Gamma_F$. Let $\hhat{H}$ and
$\hhat{M}$ denote the compactifications by adjoining $\partial F$ and
the limit set $\Lambda_F$ to $\til{H}$ and $\til{M}$ respectively.
\begin{theorem} {\bf Cannon-Thurston for Free Groups} The inclusion
$i: \til{H} \rightarrow \til{M}$ extends continuously to a map
$\hat{i}: \hhat{H} \rightarrow \hhat{M}$.
\label{ct-free}
\end{theorem}
\begin{proof} Let $p \in \til{H}$ be a base-point, and
$\lambda = [a,b]$ be a geodesic in the
intrinsic metric on $\til{H}$ and
$\lambda_h$ be the (bi-Lipschitz) hyperbolic geodesic
joining its end-points in $ \til{M}$. By Lemma \ref{contlemma} it suffices to show
that if $\lambda$ lies outside a large ball about $p$ in $\til{H}$,
then $\lambda_h$ lies outside a large ball about $p$ in
$\til{M}$.
Suppose that $\lambda$ lies outside an $n$-ball about $p$ in $\til H$, i.e. $d_{\til{H}}(\lambda, p) \geq n$.
Let $\{ D_i \}$ be a finite collection of compressing disks in $H$ such that each component
of $\partial H \setminus \bigcup_i \partial D_i$ is a pair of pants.
Since each (lift of) $D_i$ separates $\til{H}$ and since $\lambda$ lies
outside a large ball about $p$ in $\til{H}$, we conclude that there
exists such a lift $D$ lying outside an $m=m(n)$- ball about $p$ in $\til{H}$ and
that $\lambda$ lies in the component of $\til{H}\setminus D$ not containing $p$, where
$m(n) \rightarrow \infty$ as $n \rightarrow \infty$. Let $A=D\bigcup_i \sigma_i$ be the quasidisk containing $D$
constructed in Section \ref{qd},
where $\sigma_i$ is a closed curve on the lift $\til{\Sigma_i^s}$ of the $i-$th split surface $\Sigma_i^s$
to $\til M$. Also, let $\til{\Sigma_i^s} \setminus \sigma_i = {\til{\Sigma_i^s}}_+ \cup {\til{\Sigma_i^s}}_-$
where $ {\til{\Sigma_i^s}}_+ , {\til{\Sigma_i^s}}_-$ are the two components
of $\til{\Sigma_i^s} \setminus \sigma_i$. Similarly, let $\til{H}\setminus D={\til{H}}_+\cup{\til{H}}_-$,
where $ {\til{\Sigma_0^s}}_+\subset \partial {\til{H}}_+$ and
$ {\til{\Sigma_0^s}}_-\subset \partial {\til{H}}_-$. Assume without loss of generality that
$\lambda \subset {\til{H}}_+$ and $p \in {\til{H}}_-$. Let $ {\til{M}}_+= {\til{H}}_+\cup
\bigcup_i {\til{\Sigma_i^s}}_+$
and $ {\til{M}}_-= {\til{H}}_+\cup \bigcup_i {\til{\Sigma_i^s}}_-$. Also let ${\til{M}}_H=
\bigcup_i {\til{\Sigma_i^s}}$ be the union of all the horizontal split surfaces lifted to $\til M$.
Let $\alpha$ be an electro-ambient quasigeodesic joining the end-points of $\lambda$
in $ \til{M}$ and $\beta$ be an admissible quasigeodesic corresponding to $\alpha$.
Since $\beta$ is admissible, it consists of horizontal and vertical pieces.
Two cases arise:\\
a) $\beta\cap {\til{M}}_H \subset {\til{M}}_+$\\
b) $\beta\cap {\til{M}}_+\cap {\til{M}}_- \neq \emptyset$.
Roughly speaking Cases (a) and (b) correspond respectively
to the cases where $\beta$ does not or does intersect $A$ coarsely.
\noindent {\bf Case a:} $\beta\cap {\til{M}}_H \subset {\til{M}}_+$\\
By Corollary \ref{contlemma3} and Lemma \ref{contlemma4} it suffices to show that
there exists a function $\Theta : \mathbb{N} \rightarrow \mathbb{N}$ satisfying
$\Theta (n) \rightarrow \infty$ as $n \rightarrow \infty$ such that the following holds.\\
$d_{\til{H}}(\lambda , p) \geq n$ implies that $d_M(\beta_{adm}, p) \geq \Theta (n)$.
The existence of such a function $\Theta : \mathbb{N} \rightarrow \mathbb{N}$ follows exactly as in Lemma
\ref{contlemma-qd}.
\noindent {\bf Case b:} $\beta\cap {\til{M}}_+\cap {\til{M}}_- \neq \emptyset$\\
We shall say that $\beta$ crosses $A$ at $x$ if either $x \in \beta \cap A$, or if there exists
a vertical elementary admissible subpath $x \times I \subset \beta$, such that
either $(x,0) \in {\til{M}}_+$ and
$(x,1) \in {\til{M}}_-$ or $(x,1) \in {\til{M}}_+$ and
$(x,0) \in {\til{M}}_-$.
Let $r, q$ be the first and last points at which $\beta$ crosses $A$. Let $\beta_{ar},
\beta_{qb}$ be the subpaths of $\beta$ joining $a, r$ and $b, q$ respectively.
Then again, as in Case (a) above, there exists a function $\Theta : \mathbb{N} \rightarrow \mathbb{N}$ satisfying
$\Theta (n) \rightarrow \infty$ as $n \rightarrow \infty$ such that
$d_{\til{H}}(\lambda , p) \geq n$ implies that $d_M(\beta_{ar}\cup
\beta_{qb}, p) \geq \Theta (n)$. Hence by Corollary \ref{contlemma3} and Lemma \ref{contlemma4}
it follows that there exists a function $\Theta_2 : \mathbb{N} \rightarrow \mathbb{N}$ satisfying
$\Theta_2 (n) \rightarrow \infty$ as $n \rightarrow \infty$ such that
$d_{\til{H}}(\lambda , p) \geq n$ implies that $d_M(\mu_{ar}\cup
\mu_{qb}, p) \geq \Theta_2 (n)$, where $\mu_{ar}$ (resp. $\mu_{qb}$) are the (bi-Lipschitz)
hyperbolic geodesics in $\til M$ joining $a, r$ and $b, q$ respectively.
If $r, q \in A$, then by Corollary \ref{contlemma-qd1}, there exists a function $\Theta_3 : \mathbb{N} \rightarrow \mathbb{N}$ satisfying
$\Theta_3 (n) \rightarrow \infty$ as $n \rightarrow \infty$ such that
$d_{\til{H}}(\lambda , p) \geq n$ implies that $d_M(\mu_{rq}, p) \geq \Theta_3 (n)$, where $\mu_{rq}$
is the (bi-Lipschitz)
hyperbolic geodesic in $\til M$ joining $ r$ and $ q$. Hence by $\delta -$hyperbolicity
of $\til M$, $d_M(\mu_{ab}, p) \geq min(\Theta_2 (n), \Theta_3 (n)) - 2 \delta$ and we are done.
Else, let $r \in \til{\Sigma_i^s}$ and $q \in \til{\Sigma_j^s}$. There exist $C(m), m \in \mathbb{N}$
such that $d_M(r, A) \leq C(i)$ and $d_M(q, A) \leq C(j)$. Choose $r_1,
q_1$ in $\sigma_i, \sigma_j$ respectively such that $d_M(r, r_1) \leq C(i)$ and $d_M(q, q_1) \leq C(j)$.
Then, again as in the proof of Lemma \ref{contlemma-qd},
there exists a function $\Theta_4 : \mathbb{N} \rightarrow \mathbb{N}$ satisfying
$\Theta_4 (n) \rightarrow \infty$ as $n \rightarrow \infty$ such that
$d_M(\mu_{r_1q_1}, p) \geq \Theta_4 (n)$, where $\mu_{r_1q_1}$
is the (bi-Lipschitz)
hyperbolic geodesic in $\til M$ joining $ r_1$ and $ q_1$. Hence by $\delta -$hyperbolicity
of $\til M$ again, $d_M(\mu_{ab}, p) \geq min(\Theta_2 (n), \Theta_4 (n)) - 2 \delta$ and we are through.
\end{proof}
\subsection{Finitely Generated Kleinian Groups}
In this subsection, we indicate the modifications necessary in the
previous subsection to prove the analogous theorem for finitely
generated Kleinian groups.
Let $N_{gf}$ denote the augmented Scott core of $N^h
= {\Hyp}^3 / G$. Let $i: N_{gf} \rightarrow
N^h$ be the natural inclusion map. $N_{gf}$ can be naturally identified with the convex core of a geometrically finite
manifold including into $N^h$ as a type-preserving homotopy equivalence. Let $H$ be $N_{gf}$ with open neighborhoods of cusps removed.
Let $N$ be $N^h$ with open neighborhoods of cusps removed. Then $\til H$ is strongly hyperbolic relative to its horospheres and $\til N$
is strongly hyperbolic relative to its horospheres. Let $\widehat{H}$ and $\widehat{N}$ denote their relative hyperbolic compactifications. Note that
$\widehat{H}=\widehat{N_{gf}}$, where $\widehat{N_{gf}}$ is the Gromov compactification of the hyperbolic space $\widetilde{N_{gf}}$. Similarly,
$\widehat{N}=\widehat{N^h}$, where $\widehat{N^h}$ is the Gromov compactification of the hyperbolic space $\widetilde{N^h}$.
Let $\til{i}: \til{H} \rightarrow \til{N}$ indicate
the lift of $i$. Let $M$ denote the model manifold for $N$ and let $\widehat{M}$ denote the relative hyperbolic
compactification of $\til M$. We identify $H$ with its bi-Lipschitz image in $M$ under the bi-Lipschitz homeomorphism
$F:N\rightarrow M$. Let $d_G$ be the graph metric on $\til M$ equipped with a split geometry structure where all split components are incompressible.
First, suppose that
$H$ has incompressible boundary as a pared manifold. Then Theorem \ref{ptpre-inde} shows that a Cannon-Thurston map exists for $\til{i} : \til{H} \rightarrow \til{M}$. The point pre-image
description is also furnished by Theorem \ref{ptpre-inde}.
Else $H$ may be decomposed as the disk-connected sum of
$H_1$, $H_2 \cdots H_{m+1}$ for
manifolds $H_i$ where at most one of
the $H_i$'s is a handlebody without any parabolics (taken to be $H_1$ without loss of generality) and the rest
are pared manifolds with incompressible boundary. If one of the $H_i$'s (say $H_1$) is a handlebody, then we choose a maximal collection of disjoint non-separating compressing
disks in $H_1$ whose complement in
$H_1$ is a ball. Let
$D_i, i = 1 \cdots m$ denote all the compressing disks thus obtained. Since
$m \geq 1$, there is at least one compressing disk.
\begin{theorem} {\bf Cannon-Thurston for Kleinian Groups} Let $H, M, N, N_{gf}, N^h$ be as above. The inclusion
$\til{i}: \til{H} \rightarrow \til{M}$ extends continuously to a map
$\hat{i}: \hhat{H} \rightarrow \hhat{M}$ between the relative hyperbolic compactifications. Equivalently,
$\til{i}: \til{N_{gf}} \rightarrow \til{N^h}$ extends continuously to a map
$\hat{i}: \hhat{N_{gf}} \rightarrow \hhat{N^h}$ between the hyperbolic compactifications.
\label{ct}
\end{theorem}
\noindent {\bf Proof:} From Propositions
\ref{incompressible}, \ref{graph_metric_hyperbolic} and \ref{A-qc}, we
can construct quasidisks $A_i$ corresponding to $D_i$ as before and
lift them to $\til{M}$ (after partially electrocuting $\mathbb{Z}$-cusps if any).
Now, let $\lambda$ be a geodesic segment in $\til{H}$ lying outside a
large ball $B_N(p)$ for a fixed reference point $p$. $\lambda$ may be
decomposed into (at most) three pieces $\lambda_-$, $\lambda_0$ and
$\lambda_+$ as follows. \\
\begin{enumerate}
\item The middle piece
$\lambda_0$ does not intersect any of the (lifts of the) compressing
disks $D_i$ in the interior. (We thus allow for the cases where
$\lambda_-$ and/or $\lambda_+$ are empty.)
\item
the common end-point $\lambda_- \cap \lambda_0$ lies on some
$D_i$. The same is demanded of $\lambda_0 \cap \lambda_+$.
\item the point $q$ on $\lambda$ nearest to $p$ lies on $\lambda_0$
\end{enumerate}
Two cases arise.
{\bf Case A:} For a sequence of $\lambda$'s lying outside larger
and larger balls $B_n(p)$ about $p$, the last disk $D_i$ that $[p,q]$
intersects lies outside large balls $B_m(p)$ where $m \rightarrow
\infty$ as $n \rightarrow
\infty$. This is exactly the case as in the proof of Theorem
\ref{ct-free}.
The same proof goes through by Corollary \ref{contlemma-qd1}.
{\bf Case B:} There is (up to subsequencing) a fixed disk $D_i$ that is the
last disk that $[p,q]$ intersects. Since $D_i$ is of uniformly bounded
diameter, we may shift our base point to a point $p'$
in the component $\til{H_i}$ which is the lift of $H_i$ `on the other
side of' $D_i$, i.e. having $D_i$ on its boundary but not containing
$p$. In this case, there exists a fixed $n_0$ such that $\lambda$ lies
outside $B_{(n-n_0)}(p')$. By shifting origin, we rewrite $p'$ as $p$
and
$(n-n_0)$ as $n$.
\noindent{\bf Step 1:}
Now, $\lambda_0 \subset \til{H_i} \subset \til{H}$ as it does not meet any
disk $D_i$ in its interior. Since $H_i$ is either a handlebody
without parabolics, or a pared manifold with incompressible boundary, then by
Theorem \ref{ct-free} or Theorem \ref{ptpre-inde}
respectively, a Cannon-Thurston map exists for the inclusion $\til{H_i} \subset \til{M}$. By (the necessity part of) Lemma \ref{contlemma} and Lemma \ref{crit-relhyp}, it follows that the
hyperbolic geodesic $\lambda_{0h}$ joining the end-points of
$\lambda_0$ in $\til{N^h}$ lies outside a large ball about $p$. Thus,
there exists $m_1(n) \rightarrow \infty$ as $n \rightarrow \infty$
such that $\lambda_{0h}$ lies outside a ball of radius $m_1(n)$ about
$p$ in $\til{N^h}$.
\noindent{\bf Step 2:} If $\lambda_+$ (or $\lambda_-$) is non-empty,
then $\lambda_-$ (or $\lambda_+$) is separated from $p$ by a disk $D_i
\subset \til{H}$ lying outside $B_n(p)$.
Recall that $M$ is a bi-Lipschitz model for $N$, which in trun is $N^h$ with cusps removed. Also recall that $d_G$ is the graph metric on
$\til M$. Then the quasidisk $A_i$ is
quasiconvex in $(\til{M}, d_G)$ and lies outside a large ball of radius $m_2(n)$ about
$p$, where $m_2(n) \rightarrow \infty$ as $n \rightarrow \infty$. Again, by constructing admissible paths and electro-ambient
quasigeodesics as in the proof of Theorem \ref{ct-free}, we obtain a
new function $m_3(n)$ such that
$m_3(n) \rightarrow \infty$ as $n \rightarrow \infty$ and so that the
hyperbolic geodesics $\lambda_{-h}$ or $\lambda_{+h}$ lie outside a
ball of radius $m_3(N)$ about $p$ in $\til{N^h}$.
\noindent{\bf Step 3:} Therefore $\lambda_{-h} \cup \lambda_{0h} \cup
\lambda_{+h}$ lies outside a ball of radius $m_4(n) = min \{ m_1(n),
m_3(n) \}$. Finally, since $\til{N^h}$ is hyperbolic, the hyperbolic
geodesic $\lambda_h$ joining the end-points of $\lambda$ lies outside
a ball of radius $m(n) = m_4(n) - 2 \delta$ about $p$. Also,
$m(n) \rightarrow \infty$ as $n \rightarrow \infty$. Therefore, by
Lemma \ref{contlemma} or \ref{crit-relhyp}, it follows that
the inclusion
$\til{i}: \til{H} \rightarrow \til{M}$ extends continuously to a map
$\hat{i}: \hhat{H} \rightarrow \hhat{M}$ or equivalently that $\til{i}: \til{N_{gf}} \rightarrow \til{N^h}$ extends continuously to a map
$\hat{i}: \hhat{N_{gf}} \rightarrow \hhat{N^h}$. This concludes the proof. $\Box$
\begin{comment}
\medskip
\noindent {\bf Remark:} Suppose $G$ has parabolics. Then the above proof almost goes through.
The one crucial ingredient that is missing is an analogue of Theorem \ref{klar}
for pared manifolds with incompressible boundary. (Theorem \ref{klar}
is the reduction Theorem that allows results for surface groups to be pushed through for 3-manifolds with incompressible boundary.) Modulo
such a reduction Theorem, the above proof goes through verbatim.
For manifolds of bounded geometry, a direct proof (without a reduction Theorem) of the Cannon-Thurston property for pared manifolds
with incompressible boundary was given by the author in \cite{brahma-pared}. In the general
situation a direct proof without using a reduction Theorem was also given for manifolds of split geometry in \cite{brahma-amalgeo}. However
this proof is rather cumbersome and in \cite{mahan-elct2} we shall provide the necessary reduction Theorem.
\medskip
\end{comment}
Let $\hhat{{G}_F}$ denote the Floyd compactification of a group $G$
(See \cite{Floyd}). McMullen conjectured in \cite{ctm-locconn} that
there exists a continuous extension of $i:\Gamma_G \rightarrow
\til{M}$ to a map from $\hhat{G_F}$ to $\hhat{M}$. It was shown by
Floyd in \cite{Floyd} that there is a continuous map from $\hhat{G_F}$
to $\hhat{H}$. Combining this with Theorem \ref{ct} above for Kleinian groups with parabolics, we get a
proof of the following.
\begin{theorem} For any finitely generated Kleinian group $G$ and $M =
{\Hyp}^3/G$, there is a
continuous extension $\hat{i}: \hhat{G_F} \rightarrow \hhat{M}$.
\label{conj-ctm}
\end{theorem}
\section{Point pre-images of the Cannon-Thurston Map }
In this section, we determine the pre-images of points under the Cannon-Thurston map for degenerate free Kleinian groups $G$. The results are extended to arbitrary
finitely generated Kleinian groups. We shall not have need to distinguish between the hyperbolic manifold and its bi-Lipschitz model any longer and will denote
the manifold by $M$.
We set up some notation for the purposes of this section. Let $G$ be a free degenerate Kleinian group without parabolics. Suppose that $G$ is not geometrically
finite. Let $M = {\mathbb{H}}^3/G$ be the quotient manifold. Note that the limit set of $G$ is all of the sphere at infinity. Hence $M$ is its own convex core.
Let $H$ be a compact core of $M$. $H$ is a handlebody whose inclusion into $M$ induces a homotopy equivalence. In fact, $M$ deformation
retracts onto $H$. Then $\til H$ is embedded in $\til{M} = {\mathbb{H}}^3$. Let $\Gamma$ denote the Cayley graph of $G$ with respect to some finite generating set of
$G$. Assume that $\Gamma$ is embedded in $\til H$. Let $S$ denote the boundary surface of $H$. We assume that the ending lamination $\Lambda_{EL}$
is a geodesic lamination on $S$ equipped with some (any) hyperbolic metric. This is well-defined up to Dehn twists along simple closed curves in $S$ that bound
disks in $H$. To avoid this ambiguity we will refer to the ending lamination in the Masur domain as $\Lambda_{ELH}$. $M \setminus Int( H)$ is homeomorphic to
$S \times [0, \infty )$ and is bi-Lipschitz homeomorphic to an end $M_S$ of a simply degenerate hyperbolic manifold without accidental parabolics
\cite{bowditch-model} \cite{minsky-elc3}. Thus
$S \times [0, \infty ) \subset N$ equipped with its {\it intrinsic} path metric is bi-lipschitz homeomorphic to $M_S$. We shall have need to pass
interchangeably between these two below.
\subsection{EL leaves are CT leaves}
Let $i : \til H \rightarrow \til{M}$ denote the inclusion.
Let $\partial i$ denote the continuous extension of $i$
to the boundary in Theorem \ref{ct-free}. Note that the inclusion of $\Gamma$ into $\til H$ with its intrinsic metric is a quasi-isometry. So we might as
well replace the inclusion of $\Gamma$ into $\til N$ by that of $\til H$ into $\til N$. We shall show that
point pre-images under $\partial i$
correspond to
end-points of leaves of an ending lamination in the Masur domain.
The inclusion of $S$ into $H$ as its boundary induces a surjection of fundamental groups with infinitely generated kernel $N$. Let $S_N$
denote the cover of $S$ corresponding to $N$. Then $S_N \subset \til H \subset \til M$.
To distinguish between the ending lamination $\Lambda_{ELH}$ (in the Masur domain)
and bi-infinite geodesics whose end-points are identified by $\partial i$, we make the following definition.
\begin{definition}
A {\bf CT leaf} $\lambda_{CT}$ is a bi-infinite geodesic whose end-points are identified by $\partial i$. \\ An {\bf EL leaf} $\lambda_{EL}$ is a bi-infinite geodesic whose end-points
are ideal boundary points of
either a leaf of the ending lamination, or a complementary ideal polygon.
\end{definition}
We shall show that \\
$\bullet$ {\bf An {\it EL leaf} is an {\it CT leaf}.}\\
$\bullet$ {\bf A {\it CT leaf} is an {\it EL leaf}.}
\begin{prop} {\bf EL is CT } Let $G$ be a free degenerate Kleinian group without parabolics.
Let $u,v$ be either ideal
end-points of a leaf of an ending lamination of $G$, or ideal boundary points of a
complementary ideal polygon. Then
$\partial i(u) = \partial i(v)$.
\label{ptpreimage}
\end{prop}
{\bf Proof:} This is almost identical to Proposition 2.1 of \cite{mahan-elct}. However, since the setup is
somewhat different we include a proof.
Take a sequence of short geodesics $\underline{s_i}$
exiting the
end. Let $\underline{a_i}$ be geodesics in the intrinsic metric
on the boundary $S$ (of $H$)
freely homotopic to $\underline{s_i}$. By topological tameness \cite{agol-tameness} \cite{gab-cal} and geometric tameness (\cite{thurstonnotes} Ch. 9) we may assume
further that $\underline{a_i}$'s are simple closed curves on $S$. Join $\underline{a_i}$ to
$\underline{s_i}$ by the shortest geodesic $\underline{t_i}$
in $S \times [0, \infty )$ connecting the two
curves. Then the collection $\underline{a_i}$ may be chosen to converge to the ending
lamination on $S$ (\cite{thurstonnotes} Ch. 9). Also, in $S_N \subset \til H \subset \til M$, we obtain segments
$a_i \subset \til{S}$
which are finite segments whose end-points are identified by the
covering map $P: \til{S \times [0, \infty )} \rightarrow S \times [0, \infty )$. We also assume that $P$ is
injective restricted to the interior of $a_i$'s mapping to
$\underline{a_i}$.
Similarly there
exist
segments
$s_i \subset \til{M}$
which are finite segments whose end-points are identified by the
covering map $P: \til{M} \rightarrow M$. We also assume that $P$ is
injective restricted to the interior of $s_i$'s.
The finite segments $s_i$ and $a_i$ are chosen in such
a way that there exist
lifts $t_{1i}$, $t_{2i}$, joining end-points of $a_i$
to corresponding end-points of $s_i$. The union of these four pieces looks like a trapezium (see below, where we have omitted subscripts for convenience).
\begin{center}
\includegraphics[height=4cm]{trapezium.eps}
\underline{Figure: {\it Trapezium} }
\end{center}
\smallskip
Next, given any leaf $\lambda$ of the ending lamination, we may choose
translates of the finite segments $a_i$
(under the action of $\pi_1(H)$) appropriately, such that they
converge to $\lambda$ in $S_N$. For each $a_i$, let
\begin{center}
$b_i = t_{1i} \circ s_i \circ {t_{2i}^{-1}}$
\end{center}
where ${t_{2i}^{-1}}$ denotes $t_{2i}$ with orientation
reversed. If the translates of $a_i$ we are considering
have end-points lying outside large balls around a fixed reference
point $p \in S_N$, it is easy to check that $b_i$'s lie
outside large balls about $p$ in $\til{M}$. Since ${\mathbb{H}}^3$
is $\delta$-hyperbolic for some $\delta > 0$, it follows that the geodesic joining the end-points of $b_i$ (and hence $a_i$ which has the same end-points) lies in a $2\delta -$ neighborhood
of $b_i$.
At this stage we invoke the existence theorem for Cannon-Thurston
maps, Theorem \ref{ct-free}. Since $a_i$'s converge to
$\lambda$ and the hyperbolic geodesics
joining the end-points of $a_i$ exit all
compact sets, it follows that $\partial i(u) = \partial i(v)$, where
$u, v$ denote the boundary points of $\lambda$. The Proposition
follows.
$\Box$
Any finitely generated Kleinian group is geometrically tame
(\cite{agol-tameness} \cite{gab-cal} \cite{thurstonnotes} Ch. 9) and has finitely many ends.
Observe that the proof of the above Proposition used the freeness of $G$ only at the stage of applying Theorem \ref{ct-free}. The same proof goes through
verbatim for freely decomposable Kleinian groups with degenerate ends. The only modification to the above proof is that we consider one end of the manifold $M$
at a time (and the pigeon-hole principle) along with Theorem \ref{ct} in place of Theorem \ref{ct-free} to obtain the following Proposition.
\begin{prop} {\bf EL is CT - General Case} Let $G$ be a finitely generated freely decomposable Kleinian group.
Let $u, v$ be either ideal
end-points of a leaf of an ending lamination of $G$, or ideal boundary points of a
complementary ideal polygon. Then
$\partial i(u) = \partial i(v)$.
\label{ptpreimage-genl}
\end{prop}
\subsection{CT leaves are EL leaves} As usual we deal first with free
degenerate groups without parabolics.
We restate Theorem \ref{ptpre-inde} in a form that we shall use. Recall that
$M \setminus Int( H)$ is homeomorphic to
$S \times [0, \infty )$ and is bi-Lipschitz homeomorphic to an end $M_S$ of a simply degenerate hyperbolic manifold without accidental parabolics
\cite{bowditch-model} \cite{minsky-elc3}. Hence by Theorem \ref{ptpre-inde} we have the following.
\begin{theorem} \cite{mahan-elct} Let $S, M_S$ be as above.
Then the inclusion
$\tilde{j} : \widetilde{S} \rightarrow \widetilde{M_S}$ extends
continuously to the boundary. Further, pre-images of points
on the boundary are precisely ideal boundary points of a leaf of the ending
lamination $\Lambda_{ELS}$ of $M_S$, or ideal boundary points of
a complementary ideal polygon whenever the Cannon-Thurston map is not one-to-one.
\label{ptpre0}
\end{theorem}
We identify the Cayley graph $\Gamma$ of the free group with a subset of $ \til H \subset \til M$, viz. the orbit of a base-point joined by edges. The
next Thorem is one of the main Theorems of this paper.
\begin{theorem} Let $G$ be a free degenerate free Kleinian group without parabolics.
Let $i : \Gamma_G \rightarrow {\mathbb H}^3$ be the natural identification
of a Cayley graph of $G$ with the orbit of a point in ${\mathbb H}^3$. Then $i$ extends continuously to a map
$\hat{i}: \hhat{\Gamma_G} \rightarrow {\mathbb D}^3$,
where $\hhat{\Gamma_G}$ denotes the (Gromov) hyperbolic compactification of $\Gamma_G$. Let $\partial i$ denote the restriction of $\hat{i}$ to the boundary $\partial \Gamma_G$ of $\Gamma_G$.
Then $\partial i(a) = \partial i(b)$ for $a\neq b \in \partial \Gamma$ iff $a, b$ are either ideal end-points of a leaf of an ending lamination of $G$, or ideal boundary points
of a complementary ideal polygon.
\label{ptpreimagefinal}
\end{theorem}
\begin{proof}
By Theorem \ref{ct-free} the inclusion $i: \Gamma \rightarrow \til M$ extends continuously
to a map between the Gromov compactifications $\hat{i} : \widehat{\Gamma} \rightarrow {\mathbb{D}}^3$.
Let $\partial i$ denote the values of the above continuous extension to the boundary. Suppose $\partial{i} (a) = \partial{i} (b)$. $\Lambda_{EL}$ is
the ending lamination of $M$ regarded as a subset of $S$. Let $\Lambda_{ELG}$ denote $\Lambda_{EL}$ lifted to $S_G = \partial {\til H}$, which is a cover
of $S$.
We want to show
that $a, b$ are the end-points of a leaf of $\Lambda_{ELG}$. Suppose $(a,b)_\Gamma$ is the bi-infinite geodesic from $a$ to $b$ in $\Gamma
\subset \til M$. Assume without loss
of generality that $(a,b)$ passes through $1 \in \Gamma$. Let $a_k \rightarrow a$ and $b_k \rightarrow b$. Let $\overline{a_kb_k}$ denote
the geodesic in $\til M$ joining $a_k, b_k$. By continuity of the Cannon-Thurston map (Theorem \ref{ct-free}) there exists $N(k) \rightarrow \infty$
as $k \rightarrow \infty$ such that $\overline{a_kb_k}$ lies outside an $N(k)$ ball about $1 \in \Gamma \subset \til M$, where radius is measured
in the hyperbolic metric on $ \til M$. Isotoping $\overline{a_kb_k}$ slightly, we can assume without loss
of generality that it meets $\Gamma \subset \til M$ only at its end-points
(since $\Gamma$ is one dimensional). We can further isotope $\overline{a_kb_k}$ rel. endpoints by a bounded amount (depending on the Hausdorff distance
between $\til H$ and $\Gamma \subset \til{H}$) such that $\overline{a_kb_k}\cap S_G =\{ c_k, d_k \}$.\\
1) there exist $c_k, d_k \in S_G = \partial {\til H}$ with $d(a_k, c_k)$ and $d(b_k, d_k)$
uniformly bounded (independent of $k$) \\
2) if $\overline{c_kd_k}$
denotes the subpath of $\overline{a_kb_k}$ between $c_k, d_k$ then (modifying $N(k)$ by an additive constant if necessary) $\overline{c_kd_k}$
lies outside an $N(k)$ ball about $1 \in \Gamma \subset \til M$.\\
3) $\overline{c_kd_k}$ intersects $\til H$ only at the endpoints $c_k, d_k$. \\
Thus $\overline{a_kb_k}$ is a concatenation of three pieces, $\overline{a_kc_k}$, $\overline{c_kd_k}$, $\overline{d_kb_k}$, where $\overline{a_kc_k}$ and
$\overline{d_kb_k}$ are uniformly bounded in length and lie in $\til H$, whereas $\overline{c_kd_k}$ lies in $\til{M} \setminus Int(\til{H})$.
Let $[c_k,d_k]_{S_G}$ denote the geodesic in the intrinsic metric
on $S_G$ which is homotopic (rel. endpoints) to $\overline{c_kd_k}$ in $\til M \setminus Int(\til{H})$. Since $G$ is free, we can assume that its Cayley graph is a tree
and (since $\til H$ is quasi-isometric to $\Gamma$) $[c_k,d_k]_{S_G}$ passes through a point $o_k \in S_G$ at a uniformly bounded neighborhood of $1$.
Lift $[c_k,d_k]_{S_G}$ to some geodesic $[c_k,d_k] \subset \til{S} \subset \til{M_S}$ in the intrinsic metric on $\til{S}$. Further assume that there exists
some fixed $o \in \til S$ such that the corresponding lift $o_k^{\prime}$ of $o_k$ lies in a uniformly bounded neighborhood of $o$. Let
$({\overline{c_kd_k}})_{S_G}$ denote the corresponding lift of $\overline{c_kd_k}$ having the same endpoints as $[c_k,d_k]_{S_G}$ (such a choice is possible as
$[c_k,d_k]_{S_G}$ and $\overline{c_kd_k}$ are homotopic rel. endpoints in the complement of $ Int(\til{H})$ in $\til M$). It follows that
$({\overline{c_kd_k}})_{S_G}$ lies outside an $N(k)$-ball about $o_k^{\prime}$ in $\til{M_S}$. Hence (modifying $N(k)$ by a further additive constant if
necessary), $({\overline{c_kd_k}})_{S_G}$ lies outside an $N(k)$-ball about $o \in \til{M_S}$. Therefore, by the existence of Cannon-Thurston maps
for $j: \til S \rightarrow \til{M_S}$ (Theorem \ref{ptpre0}) it follows that if $[c_\infty d_\infty]_{S_G}$ denotes any subsequential
limit of the segments $[c_k,d_k]_{S_G}$ on $\til S$, then $\partial j (c_\infty) = \partial j (d_\infty) $ and hence again by
Theorem \ref{ptpre0} $c_\infty , d_\infty$ are end-points of leaves of the ending lamination $\Lambda_{ELS}$ of $S \subset M_S$. Finally, since
$(c_\infty , d_\infty )$ are bi-infinite geodesics passing through a bounded neighborhood of $o$, they project to leaves of $\Lambda_{ELG}$
in $S_G$. These leaves are also well-defined as leaves of the ending lamination $\Lambda_{ELH}$ as leaves of the ending lamination of $M$ regarded
as an element of the Masur domain. We have thus finally shown that $\Lambda_{CT} \subset \Lambda_{ELH}$. Combining this with Proposition \ref{ptpreimage}
and Theorem \ref{ct-free}
we have the Theorem.
\end{proof}
Note that in the proof of Theorem \ref{ptpreimagefinal} we have used freeness of $G$ to conclude only two things: \\
1) The manifold $M$ has exactly one end. \\
2) The path $\lambda$ in $\til H$ can be isotoped off the Cayley graph of $G$ embedded in $\til H$. \\
To prove an analogue of Theorem \ref{ptpreimagefinal} for arbitrary finitely generated Kleinian groups we continue with
the notation that $M$ is a hyperbolic manifold with augmented Scott core $H$. Then $M$
has finitely many ends. We first note, that if $\lambda =(a_\infty , b_\infty )$ is a CT leaf then there exist $a_n \rightarrow a_\infty$
and $b_n \rightarrow b_\infty$ such that the geodesic realizations $\mu_n$ of $[a_n,b_n]$ in $\til M$ leave arbitrarily large compact sets.
We may assume that $\til{M} \setminus \til{ H}$ consists of lifts of the ends of $M$ to $\til M$. If $\mu_n$ intersects more than
one such lift, it follows that there will be a subsegment $\nu_n$ of $\mu_n$ such that \\
1) $\nu_n$ is contained entirely in one of these lifts of the ends \\
2) Endpoints $c_n, d_n$ of $\nu_n$ lie on $\til H$\\
3) $c_n \rightarrow a_\infty$
and $d_n \rightarrow b_\infty$ \\
We may therefore assume without loss of generality that $\mu_n$ lies in precisely one of the lifts of the ends $E$ of $M$. If $S^h = H \cap E$ be its boundary
then the ending lamination lies in the boundary of the (relatively)
hyperbolic group $j_\ast (\pi_1(S^h))$ (hyperbolic relative to the cusp groups if any), where $j: S^h \rightarrow M$ is inclusion.
Fact (2) now goes through for arbitrary finitely generated Kleinian groups, as the inclusion of the augmented Scott core into $M$ is a
homotopy equivalence (in fact a deformation retract)
and we are only interested in leaves which are limits of segments whose geodesic realizations lie inside the lift of a fixed end.
With this modification, and with Theorem \ref{ct} in place,
the proof of Theorem \ref{ptpreimagefinal} goes through for arbitrary finitely generated Kleinian groups. However, since we have to pick one end of the manifold
at a time, the statement is a bit more involved.
\begin{theorem} Let $G$ be a finitely generated Kleinian group.
Let $i : \Gamma_G \rightarrow {\mathbb H}^3$ be the natural identification
of a Cayley graph of $G$ with the orbit of a point in ${\mathbb H}^3$. Then $i$ extends continuously to a map
$\hat{i}: \hhat{\Gamma_G} \rightarrow {\mathbb D}^3$,
where $\hhat{\Gamma_G}$ denotes the (relative) hyperbolic compactification of $\Gamma_G$. Let $\partial i$ denote the restriction of $\hat{i}$ to the boundary $\partial \Gamma_G$ of $\Gamma_G$.
Let $E$ be a degenerate end of $N^h= {\mathbb H}^3/G$ and $\til E$ a lift of $E$ to $\til{N^h}$
and let $M_{gf}$ be an augmented Scott core of $N^h$. Then the ending lamination $\LL_E$ for the end $E$ lifts to a lamination
on $\til{M_{gf}} \cap \til{E}$. Each such lift $\LL$ of the ending lamination of a degenerate end defines a relation $\RR_\LL$ on the (Gromov) Hyperbolic boundary $ \partial \widetilde{M_{gf}}$
(equal to the relative hyperbolic boundary $\partial \Gamma_G$ of $\Gamma_G$),
given by
$a\RR_\LL b$ iff $a, b$ are end-points of a leaf of $\LL$. Let $\{ \RR_i \}_i$ be the entire collection of relations on $ \partial \widetilde{M_{gf}}$ obtained this way. Let $\RR$ be the transitive closure of
the union $\bigcup_i \RR_i$. Then $\partial i(a) = \partial i(b)$ iff $a\RR b$.
\label{ptpreimagefinal-kg}
\end{theorem}
\section{Applications and Extensions}
In this section we shall first mention a couple of applications of the main Theorems of this paper. Finally we indicate an extension of the Sullivan-McMullen dictionary between complex
dynamics and Kleinian groups.
\begin{comment}
In this section we shall first describe extensions of the main Theorems of this paper to arbitrary
finitely generated Kleinian groups. Next we shall mention an application to "Primitive Stable Representations" of free groups
in $PSl_2({\mathbb C})$. Finally we indicate an extension of the Sullivan-McMullen dictionary between complex
dynamics and Kleinian groups.
\subsection{Parabolics} We have not dealt with the situation where $G$ is allowed to have parabolics in this paper. As remarked
after Theorem \ref{ct}, the {\it proof} of Theorem \ref{ct} (Existence of Cannon-Thurston maps) almost goes through.
The one crucial ingredient that is missing is an analogue of Theorem \ref{klar}
for pared manifolds with incompressible boundary. Theorem \ref{klar}
is the reduction Theorem that allows Theorems \ref{split} and \ref{ptpre} for surface groups
to be pushed through for 3-manifolds with incompressible boundary. Modulo
such a reduction Theorem, we would have a complete proof of the Cannon-Thurston property for all finitely generated Kleinian groups, i.e. we would be able
to remove the hypothesis of "no parabolics".
For manifolds of bounded geometry (with parabolics), a direct proof (without a reduction Theorem) of the Cannon-Thurston property for pared manifolds
with incompressible boundary was given by the author in \cite{brahma-pared}. In the general
situation a direct proof without using a reduction Theorem was also given for manifolds of split geometry in \cite{brahma-amalgeo}. Since all 3-manifolds
do have split geometry, this completes the proof of the Cannon-Thurston property for all finitely generated Kleinian groups.
However,
a) the proof in \cite{brahma-amalgeo} is rather cumbersome. \\
b) it seems unlikely that the relevant portion of \cite{brahma-amalgeo} will be published in this form. \\
c) Parts of the proof of this Theorem are only sketched in \cite{brahma-amalgeo}. \\
and in \cite{mahan-elct2} we shall provide the necessary reduction Theorem.\\
d) The description of point-preimages under the Cannon-Thurston map have not been described for surfaces with punctures in \cite{brahma-amalgeo}.\\
In \cite{mahan-elct2} we shall prove \\
a) A reduction Theorem along the lines of Klarreich's Theorem \cite{klarreich} for relatively hyperbolic spaces.\\
b) Describe point-preimages of Cannon-Thurston maps for punctured surface groups in terms of ending laminations. The bounded geometry case has been
resolved by Bowditch in \cite{bowditch-ct}. We will combine some of his techniques with some of ours in \cite{mahan-elct} to deduce this. \\
This will complete the programme of proving the existence of Cannon-Thurston maps for arbitrary finitely generated Kleinian groups
and describing their structure in terms of point pre-images.
\end{comment}
\subsection{Primitive Stable Representations} In \cite{minsky-primitive} Minsky
introduced and studied primitive stable representations, an open set of
$PSl_2({\mathbb C})$ characters of a nonabelian free group, on which the action of
the outer automorphism group is properly discontinuous, and which is strictly larger
than the set of discrete, faithful convex-cocompact (i.e. Schottky)
characters.
In \cite{minsky-primitive} Minsky also conjectured that \\
{\it A discrete faithful representation of $F$ is
primitive-stable if and only if every component of the ending lamination is blocking.}
Using the structure of the Cannon-Thurston map for handlebody groups, Jeon and Kim \cite{woojin} have made considerable progress towards this conjecture.
The connection between the Cannon-Thurston map for handlebody groups and primitive stable representations was indicated to the author by Yair Minsky
in a personal communication.
We briefly sketch Jeon and Kim's argument for a degenerate free Kleinian group without parabolics.
Let $F$ be a free group of rank $n$.
An element of $F$ is called primitive if it is an element of a free generating set. Let $\PP$
denote the set of bi-infinite paths $w^\ast$ obtained by iterating cyclically reduced primitive
elements.
A representation$\rho: F \rightarrow PSl_2(\mathbb{C})$ is primitive
stable if it maps elements of $\PP$ to uniform quasigeodesics in ${\mathbb{H}}^3$.
Let $\{ D_1, \cdots , D_k \}=\Delta$ be a system of disjoint non-separating compressing disks on a handlebody
$H$ cutting $H$ into a 3-ball. A free generating set of
$ F$ is dual to such a system. For a lamination $\Lambda$, the Whitehead graph
$Wh(\Lambda , \Delta)$ is defined as follows. Cut $H$ along $\Delta$ to obtain a planar
surface $\\partial H$ with $2n$ boundary components, each labeled by $D_i^+$
or $D_i^-$ . The
vertices of $Wh(\Lambda , \Delta)$ are the boundary circles of $\partial H$, with an edge whenever
two circles are joined by an arc of $\Lambda \setminus \Delta$. Jeon-Kim \cite{woojin} show following \cite{minsky-primitive} that for the ending lamination $\Lambda_E$ of a degenerate free group
without parabolics, $Wh(\Lambda_E , \Delta)$ is connected
and has no cutpoints.
Let $\rho_E$ be the associated representation.
If $\rho$ is not
primitive stable, then there exists a sequence of primitive cyclically reduced elements $w_n$ such
that $\rho(w_n^\ast)$ is not an $n-$ quasi-geodesic.
After passing to a subsequence, $w_n$ and hence $w_n^\ast$ converges to a bi-infinite geodesic $ w_\infty$ in the Cayley graph with
two distinct end points $w_+, w_-$ in the Gromov boundary of $F$. The Cannon-Thurston map identifies $w_+, w_-$ .
Hence by Theorem \ref{ptpreimagefinal} they are either the end points of
a leaf of the ending lamination or ideal points of a complementary ideal
polygon. Hence $Wh( w_\infty , \Delta)$ is connected
and has no cutpoints. It follows that $Wh( w_n , \Delta)$ is connected
and has no cutpoints for sufficiently large $n$. It follows from a Lemma due to Whitehead that
$w_n$ cannot be primitive for large $n$, a contradiction.
\subsection{Discreteness of Commensurators}
In \cite{llr} and \cite{mahan-commens}, the main results of this paper are used to prove that commensurators of finitely generated, infinite covolume, Zariski dense Kleinian
groups are discrete. The proof proceeds by showing that commensurators preserve the structure of point pre-images of Cannon-Thurston maps.
\subsection{Extending the Sullivan-McMullen Dictionary }
A celebrated theorem of Yoccoz in Complex Dynamics (see Hubbard
\cite{hubbard-yoccoz}, or Milnor
\cite{milnor-yoccoz}) proves the local connectivity of certain Julia
sets using a technique called `puzzle pieces', which consists of a
decomposition of a complex domain into pieces each of which under
iteration by a quadratic map converges to a single point. The
dynamical system can then be regarded as a semigroup ${\mathbb{Z}}_+$
of transformations acting on a complex domain.
Split components can be regarded as a 3-dimensional analogue
of puzzle pieces. Let us try to justify this analogy.
Suppose there is a group $G$ acting cocompactly on
${\mathbb{H}}^3$. Let $H \subset G$ be a subgroup. Let $G/H$ denote the coset space.
Then what we would want as the right analogue is that if one takes a sequence of elements
$g_i$ going to infinity in the
coset space, the iterates of the convex hull of the limit set of $H$ converge to a point
in the limit sphere. Thus going to infinity in the coset space $G/H$ would be the right Kleinian groups analogue of
going to infinity in the semigroup ${\mathbb{Z}}_+$
of transformations acting on a complex domain.
In the context of this paper, $H$ would be the fundamental group of a split component.
However, for us
$G$ is a surface Kleinian group and does not act co-compactly on ${\mathbb{H}}^3$. We think of the quotient space ${\mathbb{H}}^3/H$ as parametrizing
the set of normal directions to the split
component. The graph
metric gives a combinatorial distance on ${\mathbb{H}}^3/H$ and we think of $({\mathbb{H}}^3/H, d_G)$ as the analogue of the semigroup ${\mathbb{Z}}_+$.
Thus, instead of going to
infinity by iteration in the semigroup ${\mathbb{Z}}_+$, we go to infinity in the graph metric. Further,
the analogue of the requirement that iterates go to infinity, is that
the visual diameter goes to zero as we move to infinity in the graph
metric. This is ensured by hyperbolic quasiconvexity, and also
follows easily from {\bf graph quasiconvexity}. Note that {\bf graph
quasiconvexity} is a statement that gives uniform shrinking of visual
diameter to zero as one goes to infinity.
Thus we extend the Sullivan-McMullen dictionary (see
\cite{sullivan-dict}, \cite{ctm-renorm}) between Kleinian
groups and complex dynamics by suggesting the following analogy:
\begin{enumerate}
\item {\em Puzzle pieces} are analogous to {\em split components} \\
\item {\em Convergence to a point under iteration} is analogous to
{\em graph quasiconvexity} \\
\end{enumerate}
One issue that gets clarified by the above analogy is a point raised
by McMullen in \cite{ctm-locconn}. McMullen indicates that though the
Julia set $J(P_\theta )$, where
\begin{center}
$P_\theta (z) = e^{2\pi i \theta}z + z^2$
\end{center}
\noindent need not be locally connected in general by a result of Sullivan
\cite{sullivan-lc}, the limit sets of punctured torus groups are
nevertheless locally connected. Local connectivity of Julia sets would therefore not be the right analogue of
local connectivity of limit sets in this setup. Instead we look at the {\it techniques} for proving local connectivity of limit sets
vis-a-vis the {\it techniques} for proving local connectivity of Julia sets. Thus, by proposing the analogy between puzzle
pieces and split components as above, this issue is to an extent clarified. In short, the analogy is in the technique rather than in the result.
\smallskip
An analogue of the ${\mathbb{Z}}_+$ dynamical system may also be
extracted from the split geometry model. Note that each block
corresponds to a splitting of the surface group, and hence an action
on a tree. As $i \rightarrow \infty$, the split blocks $B^s_i$ and
hence the induced splittings also go to infinity, converging to a {\bf
free
action of the surface group on an ${\mathbb{R}}$-tree dual to the ending
lamination}. Thus iteration of the quadratic function corresponds to
taking a sequence of splittings of the surface group converging to a
(particular)
action on an ${\mathbb{R}}$-tree.
\smallskip
{\bf Problem:} The building of the Minsky model and its bi-Lipschitz
equivalence to a hyperbolic manifold \cite{minsky-elc1}
\cite{minsky-elc2} gives rise to a speculation that there should be a
purely combinatorial way of doing much of the work. Bowditch's
rendering \cite{bowditch-endinv}, \cite{bowditch-model}
of the Minsky, Brock-Canary-Minsky results is a step in this
direction. This paper brings out the possibility that the whole thing
should be do-able purely in terms of actions on trees. Of course there
is an action of the surface group on a tree dual to a pants
decomposition. So we do have a starting point. However, one ought to
be able to give a purely combinatorial description, {\em ab initio},
in terms of a sequence of actions of surface groups on trees
converging to an action on an ${\mathbb{R}}$-tree. This would open up the
possibility of extending these results (including those of this paper)
to other hyperbolic groups with
infinite automorphism groups, notably free groups.
|
\section{Introduction}
Consider the cubic focusing nonlinear Schr\"odinger equation in dimension three,
\begin{equation}\label{Eqn-NLS}\left\{\begin{aligned}
&iu_t+\laplacian u + u\abs{u}^2 = 0\\
&u(0,x) = u_0 : \real^{3} \to {\mathds C}.
\end{aligned}\right.\end{equation}
This is a canonical model equation arising in physics and engineering, \cite{SulemSulem}. This equation, and other closely related equations, have been the subject of many recent mathematical studies.
\begin{comment}
This equation arises as the envelope equation in the study of light traversing a dielectric Kerr medium, where the beam power affects the refractive index.
Typically, the direction of beam propagation is neglected (and taken to be time,) leading to the two-dimensional cubic NLS. In the case of an ultra-short pulse
one cannot simplify to the frame of the group velocity as it may vary along the wave packet, what is called anomalous dispersion (as in the dispersion relation).
In this case, the full structure of the underlying three-dimensional equation is required. See \cite{SulemSulem},
\cite{NewellMoloney92}
and \cite{Berge-WaveCollapseInPhysics-98} for more details.
\end{comment}
Equation (\ref{Eqn-NLS}) is locally wellposed for data $u_0\in H^s(\real^3)$, for any $s \geq \frac{1}{2}$, \cite{Cazenave03}. Higher regularity persists under local-in-time dynamics, and the maximal time $T_{max}>0$ for which $u\in C\left([0,T_{max}),H^{s}\right)$ is the same for all $s>\frac{1}{2}$,
and we have the classic blowup alternative: either $T_{max} = +\infty$ or $\norm{u(t)}_{H^{s}}\to\infty$ as $t\to T_{max}$.
\begin{comment}
Apriori, there is no reason for the existence times in $H^1$ and $H^3$ to be the same.
WRONG - see \cite[Chapter 5]{Cazenave03}.
For $s<1$ one can argue based on conservation of energy that the local wellposedness times for data in $\dot{H}^s\cap \dot{H}^1$ and $\dot{H}^s$ are the same.
There is no corresponding argument concerning $H^3$ and $H^1$, and there is no reason apriori for the local wellposedness times to be the same.
\end{comment}
Evolution by equation (\ref{Eqn-NLS}) preserves:
\begin{align}
\label{ConserveMass}
&
\int_{\real^{3}}{\abs{u(t,x)}^2\,dx}
=\int{\abs{u_0}^2\,dx} = M[u_0],
&& \text{(mass)}
\\
\label{ConserveEnergy}
&
\int{\abs{\grad_xu(t,x)}^2\,dx} - \frac{1}{2}\int{\abs{u(t,x)}^4\,dx}
= E[u(t,x)] = E[u_0],
&& \text{(energy)}
\\
\label{ConserveMoment}
&
Im\left(\int{\overline{u}(t,x)\grad u(t,x)\,dx}\right)
= Im\left(\int{\overline{u_0}\grad u_0\,dx}\right).
&& \text{(momentum)}
\end{align}
There are corresponding symmetries. If $u(t,x)$ satisfies (\ref{Eqn-NLS}), then so do the following:
\[\begin{aligned}
u(t,x+x_0)
&&& \forall\, x_0 \in \real^{3}
&& \text{(spatial translation invariance)}\\
u(t+t_0,x)
&&& \forall\, t_0\in \real
&& \text{(time translation invariance)}\\
u(t,x)e^{i\gamma_0}
&&& \forall\, \gamma_0\in\real
&& \text{(phase invariance)}\\
u(t,x-\beta_0 t)e^{i\frac{\beta_0}{2}\cdot\left(x-\frac{\beta_0}{2}t\right)}
&&& \forall\, \beta_0 \in \real^{3}
&& \text{(Galilean invariance)}\\
\lambda_0 u(\lambda_0^2t,\lambda_0 x)
&&& \forall\, \lambda_0 > 0
&& \text{(scaling invariance)}
\end{aligned}\]
Scaling invariance leaves the $\dot{H}^\frac{1}{2}(\real^3)$ norm of data unchanged and for this reason equation (\ref{Eqn-NLS}) is deemed {\it $H^{\frac{1}{2}}$-critical}.
\begin{comment}
This terminology is a primary tool for classifying nonlinear Schr\"odinger equations. The critical norm is the most regular data for which we cannot expect the length of the local wellposedness time to depend (exclusively) on the norm of the data.
\end{comment}
Equation (\ref{Eqn-NLS}) has standing wave solutions. The ansatz, $u(t,x) = e^{i t}W(x)$, leads to the elliptic PDE,
\begin{equation}\label{DefnEqn-W}\left\{\begin{aligned}
&\laplacian W - W + W\abs{W}^2 = 0,\\
&\begin{aligned}W(\abs{x}) > 0 && \text{for } x \in \real^3\end{aligned}.
\end{aligned}\right.
\end{equation}
The unique positive radial solution\footnotemark to equation (\ref{DefnEqn-W}) is the {\it ground-state} solution of equation (\ref{Eqn-NLS}). We reserve the notation $Q$ for the ground-state solution of the {\bf two-dimensional} problem,
\begin{equation}\label{DefnEqn-Q}
\left\{\begin{aligned}
&\laplacian_{\real^2} Q -Q + Q\abs{Q}^2 = 0,\\
&\begin{aligned}Q(\abs{y}) > 0 && \text{for } y \in \real^2\end{aligned}.
\end{aligned}\right.
\end{equation}
\footnotetext{
The classic proof is in \cite{Coffman-Uniqueness3DCubicGroundState-72}. For a more general proof, including other dimensions, see \cite{Weinstein-NLSSharpInterpolation-82}, \cite{BerestyckiLions-83} and \cite{Kwong-Uniqueness-89}.
For a concise overview of these results, see \cite[Appendix B]{Tao06}.
}
Recently it was shown, \cite{DHR-Scattering3DCubic-08}, that solutions to equation (\ref{Eqn-NLS}) exist for all time, and scatter, if,
\begin{equation}\label{Eqn-DHR-Result}\begin{aligned}
M[u_0]\,E[u_0] < M[W]\,E[W] \text{ and } \norm{u_0}_{L^2}\norm{\grad u_0}_{L^2} < \norm{W}_{L^2}\norm{\grad W}_{L^2}.
\end{aligned}\end{equation}
Negative energy data $u_0\in H^1$ lead to blow up in finite time if either radially symmetric or with finite variance, $u_0 \in \Sigma = H^1\cap\{f : \abs{x}f(x)\in L^2\}$, \cite{OgawaTsutsumi-NegEnerBlowupForEnergySubcrit-91}.
By adjusting the quadratic phase of negative-energy data, one can produce examples of blowup solutions with arbitrary energy, \cite[Remark 6.5.9]{Cazenave03}. Further sufficient conditions for blowup based on the virial identity are known, \cite{HolmerPlatteRoudenko-09}.
As a companion to equation (\ref{Eqn-DHR-Result}), \cite{HolmerRoudenko-SharpCondition3DCubicNLS-08} and \cite{HolmerRoudenko-DivInfVar3DCubic-10} show that if,
\[
\begin{aligned}
M[u_0]\,E[u_0] < M[W]\,E[W] \text{ and } \norm{u_0}_{L^2}\norm{\grad u_0}_{L^2} > \norm{W}_{L^2}\norm{\grad W}_{L^2},
\end{aligned}
\]
then the solution either breaks down in finite time or is unbounded in $H^1$ as $t\to\infty$.
More generally, since equation (\ref{Eqn-NLS}) is $H^1$-subcritical, local wellposedness and the scaling symmetry prove that all solutions in $H^1$ that blow up in finite time must obey the {\it scaling lower bound},
\[
\norm{u(t)}_{H^1} \geq \frac{C}{\left(T_{max}-t\right)^\frac{1}{4}}.
\]
Alternatively, the scaling lower bound can be established through energy conservation, \cite{CW-CauchyProblemHs-90}.
Numerics suggest self-similar solutions that blowup at this rate may exist \cite{SulemSulem}. Asymptotics also suggest there may be radially symmetric solutions that focus on a sphere, as that sphere collapses into a point \cite{FGW-SingularRingNumerics-2007, HolmerRoudenko-OnBlowup3DCubic-07}. The growth of $H^1$ in that case appears to be $\left(T_{max}-t\right)^{-\frac{1}{3}}$.
Recently it was shown, \cite{MR-critNormRadialL2Super-07}, that radially symmetric solutions in $H^1$ that blow up in finite time must also blowup in the critical norm, according to,
\[
\norm{u(t)}_{\dot{H}^\frac{1}{2}} \geq \abs{\log(T_{max}-t)}^C.
\]
Since equation (\ref{Eqn-NLS}) does not satisfy the pseudo-conformal symmetry, there is no explicit closed-form blowup solution based on $W$. Indeed, the present work constructs solutions of equation (\ref{Eqn-NLS}) with precise blowup rate.
Lastly, for data of finite variance $u_0\in\Sigma$, there is an integral upper bound, \cite{Merle-LimitSolutionsAtBlowup-89}, on the blowup rate,
\[\begin{aligned}
\int_0^{T_{max}}{\norm{u(t)}_{\dot{H}^1}^\mu\,dt} < +\infty
&& \text{ for }
&& 0 \leq \mu < 1.
\end{aligned}\]
\begin{comment}
Concerning the three-dimensional cubic equation, there are numerics \cite{FGW-SingularRingNumerics-2007} indicating radially symmetric collapsing sphere solutions (apparently with infinite mass) that grow like $(T_{max}-t)^{-\frac{2}{3}}$ in $H^1$.
\end{comment}
\begin{remark}[Higher Dimensions]\label{Remark-HigherDimEqn}
We will refer to the cubic nonlinear Schr\"odinger equation in dimension $N$,
\begin{equation}\label{Eqn-NLS-Ndim}
\left\{\begin{aligned}
&iu_t+\laplacian u + u\abs{u}^2 = 0\\
&u(0,x) = u_0 : \real^{N} \to {\mathds C}.
\end{aligned}\right.\end{equation}
This equation is $\dot{H}^{\frac{N}{2}-1}$-critical, and is locally wellposed for data $u_0\in H^s$, for any $s > \frac{N}{2}-1$, with the classic blowup alternative, \cite{Cazenave03}. For data $u_0 \in H^{s'}$, for some $s' > \frac{N}{2}$, higher regularity persists and, as with equation (\ref{Eqn-NLS}), the maximal time $T_{max} > 0$ for which $u\in C\left([0,T_{max}),H^s\right)$ is the same for all $s > \frac{N}{2}-1$.
\begin{comment}
In the energy-critical case, $N=4$, equation (\ref{Eqn-NLS-Ndim}) is locally wellposed for data $u_0 \in H^s$, for any $s\geq 1$, \cite{CW-RemarksNLSCriticalCase-89}. Higher regularity persists and, as with equation (\ref{Eqn-NLS}), the maximal time $T_{max}>0$ for which $u\in C\left([0,T_{max}),H^s\right)$ is the same for all $s > 1$.
Local wellposedness in the critical space is unknown in higher dimensions.
When $N>4$, equation (\ref{Eqn-NLS-Ndim}) is locally wellposed for data $u_0 \in H^s$, for any $s>\frac{N}{2}-1$. Higher regularity is only known to persist for data $u_0 \in H^s$ with $s>\frac{N}{2}$, the maximal time $T_{max}>0$ for which $u\in C\left([0,T_{max}),H^s\right)$ is the same for all $s>\frac{N}{2}$, and we have the classic blowup alternative.
\end{comment}
\end{remark}
\begin{remark}[Notation]\label{Remark-Notation}
We use
$f\lesssim g$, $f \gtrsim g$ and $f\approx g$ to denote that there exist constants $C_1, C_2>0$ such that $f \leq C_1 g$, $f \geq C_2 g$ and $C_2 g \leq f \leq C_1 g$, respectively.
Notation $f \sim g$ is used in more casual discussion to symbolize $f$ and $g$ are of the same order. We will use $\delta(\alpha)$ to denote any function of $\alpha$ with the property $\delta(\alpha) \rightarrow 0$ as $\alpha \rightarrow 0$. The exact form of $\delta$ will depend on the context. Frequently, we use the operator,
\[\begin{aligned}
\Lambda = 1 + y\cdot\grad_y, && \text{ where } y \text{ is a two-dimensional variable.}
\end{aligned}\]
Note that for $f, g\in L^2(\real^2)$ we have, $\left(\Lambda f, g\right) = -\left(f,\Lambda g\right)$.
\end{remark}
\subsection{Statement of Result}
\begin{comment}
we introduce cylindrical coordinates,
$(r,z,\theta) \in [0,\infty)\times\real\times S^{N-2}$ for $x\in\real^N$. We refer to functions that are symmetric with respect to $\theta$ as cylindrically symmetric, and we let $H^s_{cyl}(\real^N)$ denote the cylindrically symmetric subset of $H^s$.
\begin{theorem}[Higher Dimensions]
\label{Thm-MainResult-HigherD}
There exists a set of data, open in $H^{N}_{cyl}(\real^{N})$, for which the corresponding solution to equation (\ref{Eqn-NLS-Ndim})
has maximum (forward) lifetime $T_{max}<+\infty$. These solutions exhibit blowup with the log-log rate on a set of codimension two, \linebreak $\left\{x\in\real^N : (r,z) = (r_{max},z_{max})\right\}$. That is, the appropriate equivalents of (\ref{Thm-MainResult-RZ}), (\ref{Thm-MainResult-L2}), (\ref{Thm-MainResult-SingOnRing}) and (\ref{Thm-MainResult-LogLog}) hold.
\end{theorem}
\end{comment}
For all $N\geq 3$ we introduce cylindrical coordinates $x = (r,z,\theta) \in [0,\infty)\times\real\times S^{N-2}$ for $x\in\real^N$. We refer to functions that are symmetric with respect to $\theta$ as cylindrically symmetric, and we let $H^s_{cyl}(\real^N)$ denote the cylindrically symmetric subset of $H^s$.
\begin{theorem}[Main Result]\label{Thm-MainResult}
For all $N\geq 3$, there exists a set of cylindrically symmetric data $u_0 \in {\mathcal P}$, open in $H^{N}_{cyl}(\real^{N})$ for which the corresponding solution $u(t)$ of (\ref{Eqn-NLS}) has maximum (forward) lifetime $0 < T_{max} < +\infty$ and exhibits the following properties:
\begin{itemize}
\item \underline{\rm Concentration}:
There exist parameters $\lambda(t)>0$, $r(t) > 0$, $z(t) \in \real^{N-2}$ and $\gamma(t) \in \real$, with convergence,
\begin{equation}\label{Thm-MainResult-RZ}\begin{aligned}
\left(r(t),z(t)\right) \longrightarrow \left(r_{max},z_{max}\right)
&& \text{ as } t \rightarrow T_{max}
&& \text{ with } r_{max} \sim 1,
\end{aligned}\end{equation}
such that there is the following strong convergence in $L^2(\real^{N})$,
\begin{equation}\label{Thm-MainResult-L2}\begin{aligned}
u(t,r,z,\theta) - \frac{1}{\lambda(t)}Q\left(\frac{(r,z)-(r(t),z(t))}{\lambda(t)}\right)e^{-i\gamma(t)}
\longrightarrow
u^*(r,z,\theta),
&& \text{ as } t \rightarrow T_{max}.
\end{aligned}\end{equation}
\item \underline{\rm Persistent regularity away from singular ring}:
For any $R > 0$,
\begin{equation}\label{Thm-MainResult-SingOnRing}
u^* \in H^\frac{N-1}{2}\left(\abs{(r,z)-(r_{max},z_{max})}>R\right).
\end{equation}
\item \underline{\rm Log-log blowup rate}:
The solution leaves $H^1$ at the log-log rate,
\begin{equation}\label{Thm-MainResult-LogLog}\begin{aligned}
\frac{
\left(\frac{\log\abs{\log T_{max}-t}}{T_{max}-t}\right)^\frac{1}{2}
}{
\norm{u(t)}_{H^1(\real^{3})}
}
\longrightarrow \frac{\sqrt{2\pi}}{\norm{Q}_{L^2(\real^2)}}
&& \text{ as } t \rightarrow T_{max}.
\end{aligned}\end{equation}
\begin{comment}
\lambda(t)\,\left(\frac{\log\abs{\log(T_{max}-t)}}{T_{max}-t}\right)^\frac{1}{2}
\longrightarrow \frac{\sqrt{2\pi}}{\norm{Q}_{L^(\real^2)}}
&& \text{ as } t \rightarrow T_{max},
\end{comment}
Moreover, the higher-order norm behaves appropriately,
\begin{equation}\label{Thm-MainResult-LogLogHigher}\begin{aligned}
\frac{
\norm{u(t)}_{H^{N}}
}{
\norm{u(t)}_{H^1}^{N}\,\log\norm{u(t)}_{H^1}
}
\longrightarrow 0
&& \text{ as } t \rightarrow T_{max}.
\end{aligned}\end{equation}
\end{itemize}
\end{theorem}
\begin{remark}[Proof for $N>3$]
To simplify the exposition, we only present proof in the case $N=3$.
The adaptations for higher dimensions are indicated by Remarks \ref{Remark-HigherDimEqn}, \ref{Remark-HigherDim-Hypo}, \ref{Remark-HigherDim-StrategyUseHCrit} and \ref{Remark-HigherDim-NthEnergy}.
\end{remark}
\begin{remark}[Nature of $u^*$]
\label{Remark-ProfileH1Away}
For the $L^2$-critical problem it is known that the residual profile $u^*$ is not in $H^1$, \cite{MR-ProfilesQuantization-05}. Indeed, equation (\ref{Thm-MainResult-SingOnRing}) fails for $R=0$. See Remark \ref{Remark-ProfileNonH1} for further comment.
\end{remark}
\begin{comment}
\begin{remark}[Blowup Set]
\label{Remark-SingOnRingIsBlowupSet}
The local wellposedness of (\ref{Eqn-NLS}) in $H^1(\real^3)$ may only fail at time $t=T_{max}$ due to $H^1 \rightarrow \infty$.
Therefore, (\ref{Thm-MainResult-L2}) and (\ref{Thm-MainResult-SingOnRing}) give alternate justification to view the ring $(r_{max},z_{max},\theta)$ as the singular set.
\end{remark}
\end{comment}
\subsection{Brief Heuristic}
In cylindrical coordinates we write the Laplacian,
\begin{equation}\label{Eqn-Laplacian}
\laplacian_x = \partial_r^2 + \partial_z^2 + \frac{\partial_r}{r}.
\end{equation}
Suppose that a solution to equation (\ref{Eqn-NLS}) is cylindrically symmetric and concentrated near the ring $(r,z) \sim (r_0,z_0)$. Then for an appropriately small $\lambda_0 > 0$ we may write,
\begin{equation}\label{Eqn-BriefHeuristic1}
u(t,x) = \frac{1}{\lambda_0}v\left(\frac{t}{\lambda_0^2},\frac{(r,z)-(r_0,z_0)}{\lambda_0}\right),
\end{equation}
where the function $v$ is supported on the half-plane $(r,z)\in \left\lbrack -\frac{r_0}{\lambda_0},\infty\right) \times \real$.
Neglect that our parameters may vary in time. After changing coordinates, $v$ satisfies,
\begin{equation}\label{Eqn-BriefHeuristic2}
\begin{aligned}
i\partial_{s}v + \laplacian_y v
+ \frac{\lambda_0}{r}\partial_{y_1} v
+ v\abs{v}^2 =0
&&\text{ where }s=\frac{t}{\lambda_0^2},
&& y = \frac{(r,z)-(r_0,z_0)}{\lambda_0}.
\end{aligned}
\end{equation}
For a solution $u(t,x)$ tightly concentrated near $(r_0,z_0)$, we might choose $\lambda_0 \ll 1$ as the width of the window of concentration. Then, $\frac{\lambda_0}{r}\partial_{y_1}v$ can be taken as a lower order correction, and the evolution of $v$ is essentially that of two-dimensional cubic NLS. If $v(s,y)$ falls within the robust log-log blowup dynamic, we would expect the concentration near $(r_0,z_0)$ to increase, and for the lower order correction in equation (\ref{Eqn-BriefHeuristic2}) to become less relevant.
We may identify our main challenge: to ensure persistence of sufficient decay in the original variables near $r=0$ such that conditions there mimic those at infinity during a log-log blowup of two-dimensional cubic NLS.
\subsection{Similar Results}
\begin{theorem}[Standing Ring Blowups for Quintic NLS in 2D \cite{R06}]\label{Thm-Raph}
There exists a set of radially symmetric data $u_0$, open in $H^{1}_{rad}(\real^2)$, for which the corresponding solution to,
\begin{equation}\label{Eqn-QuinticNLS2D}
\left\{\begin{aligned}
&iu_t+\laplacian u + u\abs{u}^4 = 0,\\
&u_0 \in H^1(\real^2),\end{aligned}\right.
\end{equation}
has maximum lifetime $T_{max}<+\infty$ and exhibits log-log blowup on a ring of fixed radius; that is, there holds appropriate equivalents of (\ref{Thm-MainResult-RZ}), (\ref{Thm-MainResult-L2}), (\ref{Thm-MainResult-SingOnRing}) and (\ref{Thm-MainResult-LogLog}).
\end{theorem}
The following result and Theorem \ref{Thm-MainResult} were developed simultaneously.\footnotemark
\footnotetext{After Theorem \ref{Thm-Raph}, the idea to consider other $H^\frac{1}{2}$-critical problems was first suggested to the author's thesis advisor by Justin Holmer and Svetlana Roudenko in private conversation. }
\begin{theorem}[Standing Ring Blowups for Cubic NLS in 3D \cite{HolmerRoudenko-3DCubicCircle-10}]\label{Thm-HolmerRoudenko}
There exists a set of cylindrically symmetric data $u_0$, open in $H^1_{cyl}(\real^3)$, for which the corresponding solution to (\ref{Eqn-NLS}) has maximum lifetime $T_{max} < +\infty$ and exhibits log-log blowup on a ring of fixed radius. In particular, (\ref{Thm-MainResult-RZ}) and (\ref{Thm-MainResult-L2}) hold, as does (\ref{Thm-MainResult-SingOnRing}) at the level of $H^\frac{1}{2}$ regularity.
\end{theorem}
The methods of \cite{R06} do not extend to prove either Theorem \ref{Thm-MainResult} or Theorem \ref{Thm-HolmerRoudenko}. At issue is the initial localized gain of regularity, \cite[equation (4.137)]{R06}. Calculating $\frac{d}{dt}\norm{D^\nu\left(\chi u\right)}_{L^2}^2$ results in a nonlocal term, due to the lack of Leibniz rule - see (\ref{Strategy2-KatoStrich}). In this place \cite{R06} relies upon the Strauss radial embedding, and \cite{HolmerRoudenko-3DCubicCircle-10} use an elegant microlocal estimate to smooth the nonlocal part. We will avoid such problems through the use of higher regularity.
Note that Theorem \ref{Thm-HolmerRoudenko} describes a larger class of data, with lower regularity, than does Theorem \ref{Thm-MainResult} in the case $N=3$.
\begin{theorem}[Standing Ring Blowups for Quintic NLS \cite{RaphaelSzeftel-StandingRingNDimQuintic-08}]\label{Thm-RaphSzef}
For all $N\geq 3$, there exists a set of radially-symmetric data $u_0$, open in $H^{N}_{rad}(\real^{N})$, for which the corresponding solution to:
\begin{equation}\label{Eqn-QuinticNLS}
\left\{\begin{aligned}
&iu_t + \laplacian u + u\abs{u}^4 = 0,\\
&u_0 \in H^{N}(\real^{N}),\end{aligned}\right.
\end{equation}
has maximum lifetime $T_{max}<+\infty$ and exhibits log-log blowup on a fixed ring $r\sim 1$; that is, there holds appropriate equivalents of (\ref{Thm-MainResult-L2}), (\ref{Thm-MainResult-RZ}), (\ref{Thm-MainResult-LogLog}) and (\ref{Thm-MainResult-SingOnRing}).
\end{theorem}
It is essential to the proof of both Theorem \ref{Thm-MainResult} and Theorem \ref{Thm-RaphSzef} that the behaviour of a higher-order norm can be controlled in terms of the (understood) behaviour of the $H^1$ norm. Again, the fundamental obstruction is $H^1 \not\hookrightarrow L^\infty$, which manifests in two ways:
\begin{enumerate}
\item Inability to bootstrap a global $H^3$ control. Without radial symmetry, one cannot make the estimate, $\frac{d}{dt}\norm{u}_{H^3}^2 \lesssim \norm{u}_{H^3}^{2-\delta}\norm{u}_{H^1}^{3\delta+2}$ - see the third term of \cite[equation (44)]{RaphaelSzeftel-StandingRingNDimQuintic-08}.
\item Inability to achieve an initial localized gain of regularity.
As with the arguments of \cite{R06}, an analogue of \cite[equation (63)]{RaphaelSzeftel-StandingRingNDimQuintic-08} cannot be established. To bypass this issue, we require tighter control of higher regularity than is achieved in \cite{RaphaelSzeftel-StandingRingNDimQuintic-08}, and a Brezis-Gallou\"et type argument, Lemma \ref{Lemma-NdimEndpointSobolev}.
\end{enumerate}
We apply three new strategies:
\begin{enumerate}
\item Consider the singular and residual portions of the solution separately during the most difficult calculations.
This is recognition that the higher order norms of the central profile of the solution scale exactly with the global $H^1$ norm. While the worst hypothesized $H^3$ behaviour may only be attributed to the residual portion of the solution, for the same portion we possess superior $H^1$ control. We arrange to only evaluate the $H^3$ norm of the residual when there is also a factor of the $H^1$ norm of the same.
By making a delicate hypothesis of the global $H^3$ behaviour, only slightly worse than scaling, we can arrange for product of the $H^1$ and $H^3$ norms of the residual to be slightly better than scaling.
\item Integrate the modulation parameters with more accuracy to allow a more refined hypothesis of the global $H^3$ behaviour. See Lemma \ref{Lemma-lambdaIntegral}.
\item Apply a modified Brezis-Gallou\"et argument to the residual portion of the solution to estimate $\norm{u}_{L^\infty}$ much closer to scaling than any two-dimensional Sobolev embedding. This allows us to complete a Gronwall argument and prove a localized gain of regularity outside the support of some bootstrap hypotheses.
\end{enumerate}
\begin{remark}[Theorem \ref{Thm-MainResult} for other nonlinearities]
Other $H^\frac{1}{2}$-critical equations, such as $\textrm{NLS}^{1+\frac{4}{3}}(\real^4)$ and $\textrm{NLS}^{2}(\real^5)$ suffer from non-smooth nonlinearities. The author is unaware of local wellposedness in any space more regular than $H^1$ in these cases. Non-smooth nonlinearities also prohibit iteration of interior regularity arguments, so that in these cases Lemma \ref{Lemma-UnprovenProperty} below is only true for $s=3$.
\end{remark}
\begin{comment}
For all $N\geq 3$ we introduce cylindrical coordinates,
$(r,z,\theta) \in [0,\infty)\times\real\times S^{N-2}$ for $x\in\real^N$. We refer to functions that are symmetric with respect to $\theta$ as cylindrically symmetric, and we let $H^s_{cyl}(\real^N)$ denote the cylindrically symmetric subset of $H^s$.
\begin{theorem}[Higher Dimensions]
\label{Thm-MainResult-HigherD}
There exists a set of data, open in $H^{N}_{cyl}(\real^{N})$, for which the corresponding solution to equation (\ref{Eqn-NLS-Ndim})
has maximum (forward) lifetime $T_{max}<+\infty$. These solutions exhibit blowup with the log-log rate on a set of codimension two, \linebreak $\left\{x\in\real^N : (r,z) = (r_{max},z_{max})\right\}$. That is, the appropriate equivalents of (\ref{Thm-MainResult-RZ}), (\ref{Thm-MainResult-L2}), (\ref{Thm-MainResult-SingOnRing}) and (\ref{Thm-MainResult-LogLog}) hold.
\end{theorem}
\noindent To simplify the exposition, we only present proof in the case $N=3$.
Remarks \ref{Remark-HigherDim-Hypo}, \ref{Remark-HigherDim-StrategyUseHCrit} and \ref{Remark-HigherDim-NthEnergy} will indicate how to adapt the proof.
\end{comment}
\subsection{Acknowledgements}
\label{Subsec-Ack}
The author would like to thank Pierre Rapha\"el and J\'er\'emie Szeftel for correspondence and for sharing the draft of their work, \cite{RaphaelSzeftel-StandingRingNDimQuintic-08}, and Justin Holmer and Svetlana Roudenko for conversations. The author is also indebted to his advisor, Jim Colliander, for his attention and guidance. This work is part of the author's Ph.D. thesis.
\section{Setting of the Bootstrap}
\label{Section-SetBootstrap}
In this chapter we identify data concentrated near the ring $(r,z)\sim (1,0)$, according to properties we will later show persist. Our subsequent arguments are based on the two-dimensional $L^2$-critical log-log blowup dynamic, which has been comprehensively investigated by Merle \& Rapha\"el, \cite{MR-BlowupDynamic-05, MR-SharpUpperL2Critical-03, MR-UniversalityBlowupL2Critical-04, R-StabilityOfLogLog-05, MR-ProfilesQuantization-05, MR-SharpLowerL2Critical-06}. This work stems from those detailed studies.
\begin{definition}[Fundamental Properties of Almost Self-similar Profiles]\label{Defn-FundSelfSimProps}
For all $b> 0$ sufficiently small, there exists a solution $\widetilde{Q}_b \in H^1(\real^2)$ of,
\[
\laplacian \widetilde{Q}_b -\widetilde{Q}_b + ib\Lambda\widetilde{Q}_b + \widetilde{Q}_b\abs{\widetilde{Q}_b}^2 = -\Psi_b,
\]
that is supported on the ball of radius $\frac{2}{\abs{b}}$ and converges to $Q$ in $C^3(\real^2)$ as $b\rightarrow 0$. Profiles $\widetilde{Q}_b$ have mass the order of $b^2$ larger than $Q$, and energy of the order $e^{-\frac{C}{b}}$. The truncation error $\Psi_b$ acts as the source of the linear radiation,
\[
\laplacian \zeta_b - \zeta_b + ib\Lambda\zeta_b = \Psi_b.
\]
Radiation $\zeta_b$ is not in $L^2$, with the precise decay rate $\Gamma_b = \lim_{\abs{y} \to +\infty}\abs{y}\abs{\zeta_b}^2$. It is known $\Gamma_b \sim e^{-\frac{\pi}{b}}$, and it is this decay property linked to the central profile $\widetilde{Q}_b$ that is
responsible for the log-log rate of the two-dimensional $L^2$-critical problem. For our analysis, we will truncate $\zeta_b$ near $\abs{y} \sim e^{+\frac{a}{b}}$ for a small fixed parameter $a$. See Section \ref{Subsec-SelfSim} for details.
\end{definition}
\begin{lemma}[Smoothness of $\widetilde{Q}_b$]
\label{Lemma-UnprovenProperty}
The almost self-similar profiles $\widetilde{Q}_b$ are smooth. For any $s\geq 3$,
\begin{equation}\label{Eqn-QbDerivProperty}\begin{aligned}
\limsup_{b\to 0} \norm{\widetilde{Q}_b}_{C^{s}(\real^2)} < +\infty
&& \text{ and }
&& \limsup_{b\to 0}\norm{\widetilde{Q}_b}_{H^{s}(\real^2)} < +\infty.
\end{aligned}\end{equation}
\end{lemma}
\subsection{Geometric Decomposition}
\label{Subsec-GeoDecomp}
In place of $(r,z,\theta) \in \real^{3}$ we change coordinates to the rescaled half-plane,
\begin{equation}\label{DefnEqn-y}
y = \left(\frac{(r,z) - (r_0,z_0)}{\lambda_0}\right) \in \left\lbrack-\frac{r_0}{\lambda_0},+\infty\right)\times\real.
\end{equation}
Fixed parameters $r_0$,$z_0$,$\lambda_0$ will later be replaced by $r(t)$, $z(t)$, and $\lambda(t)$. It will be clear from the context. Note the measure due to cylindrical symmetry, $dx = \lambda_0 \mu_{\lambda_0,r_0}(y)\,dy$ is given by,
\begin{equation}\label{DefnEqn-mu}
\mu_{\lambda_0,r_0}(y) = 2\pi\left(\lambda_0y_1+r_0\right){\mathds 1}_{y_1 \geq -\frac{r_0}{\lambda_0}}.
\end{equation}
We will shortly hypothesize parameters of the decomposition in such a way that the support of both $\widetilde{Q}_b$ and $\widetilde{\zeta}_b$ are well away from the boundary of domain (\ref{DefnEqn-y}). For convenience we will omit the constant factor $2\pi$ and approximate $\mu(y)\sim 1$ on this region - see (\ref{Hypo1-consequenceForMu}). Integrals in $y$ can then be seen as taken over all of $\real^2$, and regular integration by parts applies.
Any integral that cannot be localized in this way will be treated separately, and very carefully.
To begin, we modulate suitable cylindrically symmetric data as if two-dimensional, \cite[Lemma 2]{R06}.
\begin{lemma}[Existence of Geometric Decomposition at a Fixed Time]\label{Lemma-GeoDecompFixed}
Suppose that $v\in H^1_{cyl}(\real^{3})$ may be written in the form,
\begin{equation}\label{Lemma-GeoDecompFixed-Eqn}
v(r,z,\theta) =
\frac{1}{\lambda_v}\left(\widetilde{Q}_{b_v}+\epsilon_v\right)
\left(\frac{(r,z)-(r_v,z_v)}{\lambda}\right)e^{-i\gamma_v}
\end{equation}
for some parameters $\lambda_v,b_v,r_v > 0$ and $\gamma_v,z_v \in \real$ such that,
\begin{equation}\label{Lemma-GeoDecompFixed-eps}
\int{\abs{\grad_y\epsilon_v}^2\mu_{\lambda_v,r_v}(y)\,dy}
+\int_{\abs{y}\leq\frac{10}{b_v}}{\abs{\epsilon_v}^2e^{-\abs{y}}\,dy}
< \Gamma_{b_v}^\frac{1}{2},
\end{equation}
\begin{equation}\label{Lemma-GeoDecompFixed-params}\begin{aligned}
\abs{(r_v,z_v) - (1,0)} < \frac{1}{3}
&& \text{ and }
&& 10\lambda_v < b_v < \alpha^*.
\end{aligned}\end{equation}
Then there are nearby parameters: $\lambda_0,b_0,r_0 > 0$ and $\gamma_0,z_0 \in \real$ with,
\begin{equation}\label{Lemma-GeoDecompFixed-nearby}
\abs{b_0-b_v}
+\abs{\frac{\lambda_0}{\lambda_v} - 1}
+\frac{\abs{(r_0,z_0)-(r_v,z_v)}}{\lambda_v}
\leq \Gamma_{b_0}^\frac{1}{5},
\end{equation}
such that the corresponding $\epsilon_0$,
\begin{equation}\label{Lemma-GeoDecompFixed-defnEps}
\epsilon_0(y) = \lambda_0\,v\left(\lambda_0 y + (r_0,z_0)\right)\,e^{i\gamma_0} - \widetilde{Q}_{b_0},
\end{equation}
satisfies the two-dimensional orthogonality conditions\footnotemark:
\begin{equation}\label{Lemma-GeoDecompFixed-orthog}
Re\left(\epsilon_0,\abs{y}^2\widetilde{Q}_{b_0}\right) =
Re\left(\epsilon_0,y\widetilde{Q}_{b_0}\right) =
Im\left(\epsilon_0,\Lambda^2\widetilde{Q}_{b_0}\right) =
Im\left(\epsilon_0,\Lambda\widetilde{Q}_{b_0}\right) = 0.
\end{equation}
\end{lemma}
\footnotetext{
The decomposition of \cite{MR-SharpUpperL2Critical-03} used slightly different orthogonality conditions. Equation (\ref{Lemma-GeoDecompFixed-orthog}) is the decomposition introduced \cite[Lemma 6]{MR-UniversalityBlowupL2Critical-04}, which leads to a better estimate on the phase parameter than was achieved in \cite{MR-SharpUpperL2Critical-03}.
}
We envisage a singular ring contained within a toroid, the complement of which is contiguous and includes both the origin and infinity. Denote two smooth cutoff functions,
\begin{equation}\label{DefnEqn-Chi}\begin{aligned}
\chi(r,z,\theta) = &\left\{\begin{aligned}
1 && \text{ for } \abs{(r,z) - (1,0)} \geq \frac{2}{3}\\
0 && \text{ for } \abs{(r,z) - (1,0)} \leq \frac{1}{3}
\end{aligned}\right.
&& \text{and},\\
\chi_0(r,z,\theta) = &\left\{\begin{aligned}
1 && \text{ for } \abs{(r,z) - (1,0)} \geq \frac{1}{7}\\
0 && \text{ for } \abs{(r,z) - (1,0)} \leq \frac{1}{8}.
\end{aligned}\right.
&&
\end{aligned}\end{equation}
In Chapter \ref{Section-BootAtInfty} we will define a further series of cutoff functions $\psi$ and $\varphi$; these further definitions will be supported on bounded regions where $\chi_0 \equiv 1$.
We now describe the initial data for our bootstrap procedure.
\begin{definition}[Description of Initial Data ${\mathcal P}$]\label{Defn-DataP}
For $\alpha^* > 0$, a constant to be determined, let the set ${\mathcal P}(\alpha^*)$ comprise cylindrically symmetric $u_0 \in H^{3}_{cyl}(\real^{3})$ that may be written of the form,
\begin{equation}\label{Defn-DataP1-GeoDecomp}\begin{aligned}
u_0(r,z) &= \frac{1}{\lambda_0}\left(\widetilde{Q}_{b_0}+\epsilon_0\right)
\left(\frac{(r,z) - (r_0,z_0)}{\lambda_0}\right)e^{-i\gamma_0}\\
&=\frac{1}{\lambda_0}\left(\widetilde{Q}_{b_0}\right)
\left(\frac{(r,z) - (r_0,z_0)}{\lambda_0}\right)e^{-i\gamma_0}
+ \tilde{u}_0(r,z),
\end{aligned}\end{equation}
in a way that satisfies the following two sets of conditions:
\begin{description}
\item[{Singularity of a log-log nature:}]
\item[C1.1] {\it 'Radial' profile focused near a singular ring,}
\begin{equation}\label{DataP1-rz}
\abs{(r_0,z_0) - (1,0)} < \alpha^*.
\end{equation}
\item[C1.2] {\it 'Radial' profile is close to $Q$ near the singular ring,}
Profile $\widetilde{Q}_b$ have nearly the mass of $Q$, and account for nearly all mass globally,
\begin{equation}\label{DataP1-mass}\begin{aligned}
0 < b_0 + \norm{\tilde{u}_0}_{L^2(\real^{3})} < \alpha^*,
\end{aligned}\end{equation}
and $\epsilon_0(y)$ both satisfies the orthogonality conditions,
\begin{equation}\label{DataP1-orthog}\begin{aligned}
Re\left(\epsilon_0,\abs{y}^2\widetilde{Q}_{b_0}\right)
= Re\left(\epsilon_0,y\widetilde{Q}_{b_0}\right)
= Im\left(\epsilon_0,\Lambda^2\widetilde{Q}_{b_0}\right)
= Im\left(\epsilon_0,\Lambda\widetilde{Q}_{b_0}\right) = 0,
\end{aligned}\end{equation}
and the smallness condition,
\begin{equation}\label{DataP1-epsilon}
\int{ \abs{\grad_y\epsilon_0(y)}^2 \mu_{\lambda_0,r_0}(y)\,dy}
+\int_{\abs{y}\leq\frac{10}{b_0}}{\abs{\epsilon_0(y)}^2e^{-\abs{y}}\,dy}
<\Gamma_{b_0}^\frac{6}{7}.
\end{equation}
\item[C1.3] {\it Conformal and scaling parameters are consistent with log-log blowup speed,}
\begin{equation}\label{DataP1-loglog}
e^{-e^\frac{2\pi}{b_0}} < \lambda_0 < e^{-e^{\frac{\pi}{2}\frac{1}{b_0}}}.
\end{equation}
\item[C1.4] {\it Energy and localized momentum are normalized,}
\begin{equation}\label{DataP1-enerMoment}
\lambda_0^2\abs{E_0}
+\lambda_0\abs{Im\left( \int{\grad_x\psi^{(x)}\cdot\grad_xu_0\overline{u}_0}\right)}
< \Gamma_{b_0}^{10},
\end{equation}
where $\psi^{(x)}$ is a smooth cylindrically symmetric `cutoff' function with,
\begin{equation}\label{DefnEqn-psi-x}
\psi^{(x)}(r,z,\theta) = \left\{\begin{aligned}
r + z && \text{ for } \abs{(r,z)-(1,0)}\leq\frac{1}{2},\\
0 && \text{ for } \abs{(r,z)-(1,0)}\geq\frac{3}{4}.
\end{aligned}\right.
\end{equation}
\item[{Regularity away from the singularity:}]
\item[C2.1] {\it Scaling-consistent $\dot{H}^{3}$ norm,}
\begin{equation}\label{DataP2-Hnk}
\norm{u_0}_{H^{3}(\real^{3})} < \frac{C_{\widetilde{Q}}}{\lambda_0^{2+k}},
\end{equation}
where $C_{\widetilde{Q}}$ is a universal constant due to Lemma \ref{Lemma-UnprovenProperty},
\item[C2.2] {\it Strong hierarchy of regularity away from the singular ring,}
\begin{equation}\label{DataP2-Hlower}
\norm{\chi_0u_0}_{H^{3-\kappa}(\real^{3})} < \frac{1}{\lambda_0^{3-2\kappa}},
\end{equation}
for each half integer $\frac{1}{2} \leq \kappa \leq \frac{3}{2}$, and,
\item[C2.3] {\it Vanishing lower-order norms away from the singular ring,}
\begin{equation}\label{DataP2-Hcrit}
\norm{\chi_0u_0}_{H^1(\real^{3})} < \left(\alpha^*\right)^\frac{1}{2}.
\end{equation}
\end{description}
\end{definition}
Lemma \ref{Lemma-GeoDecompFixed} guarantees that ${\mathcal P}(\alpha^*)$ is open in $H^1_{cyl}\cap H^3_{cyl}$. See Appendix \ref{Appendix}
for proof that ${\mathcal P}(\alpha^*)$ is nonempty.
For the remainder of this paper, fix an arbitrary $u_0 \in {\mathcal P}(\alpha^*)$. Let $u(t)$ denote the evolution by equation (\ref{Eqn-NLS}), with maximum (forward) lifetime, $0 < T_{max} \leq +\infty$.
Continuous evolution in $H^3(\real^3)$ implies the same in $H^1(\real^3)$, and so by Lemma \ref{Lemma-GeoDecompFixed} there is some $0<T_{geo} \leq T_{max}$, (which may be assumed maximal,) for which the geometric decomposition of Lemma \ref{Lemma-GeoDecompFixed} may be applied on $[0,T_{geo})$.
There exist unique continuous functions $\lambda(t),b(t),r(t) : [0,T_{geo}) \to (0,\infty)$, and $\gamma(t),z(t) : [0,T_{geo}) \to \real$, with the expected initial values, where,
\begin{equation}\label{Eqn-GeoDecomp}\begin{aligned}
u(t,r,z,\theta) =
&\frac{1}{\lambda(t)}\left(\widetilde{Q}_{b(t)}+\epsilon(t)\right)
\left(\frac{(r,z)-(r(t),z(t))}{\lambda}\right)e^{-i\gamma(t)}\\
=&\frac{1}{\lambda(t)}\left(\widetilde{Q}_{b(t)}\right)
\left(\frac{(r,z)-(r(t),z(t))}{\lambda}\right)e^{-i\gamma(t)}
+\tilde{u}(t,r,z,\theta),
\end{aligned}\end{equation}
such that $\epsilon(t,y)$ satisfies the two-dimensional orthogonality conditions:
\begin{align}
&&Re\left(\epsilon(t),\abs{y}^2\widetilde{Q}_{b(t)}\right)
&= 0,\label{Eqn-GeoDecomp-OrthogReY2}\\
&&Re\left(\epsilon(t),y\widetilde{Q}_{b(t)}\right)
&= 0,\label{Eqn-GeoDecomp-OrthogReY}\\
&&Im\left(\epsilon(t),\Lambda^2\widetilde{Q}_{b(t)}\right)
&= 0,\label{Eqn-GeoDecomp-OrthogImL2}\\
&\text{and}&Im\left(\epsilon(t),\Lambda\widetilde{Q}_{b(t)}\right)
&= 0.\label{Eqn-GeoDecomp-OrthogImL}
\end{align}
We may now define the rescaled time,
\begin{equation}\label{DefnEqn-s}\begin{aligned}
s(t) = \int_0^t{\frac{1}{\lambda^2(\tau)}d\,\tau} + s_0
&& \text{ where }
&& s_0 = e^\frac{3\pi}{4b_0}.
\end{aligned}\end{equation}
Also denote $s_1 = s(T_{hyp})$, for $T_{hyp}$ due to the forthcoming Definition \ref{Defn-Hypo}.
The choice of $s_0$ will prove convenient in Section \ref{Subsec-LocalVirial}.
\begin{remark}[Fixed Parameters]
To aid the reader, we provide a brief summary of the various parameters that will be introduced, in the order one might ultimately determine them:
\begin{itemize}
\item $\eta$ and $a$: Parameters that determine the cutoff shape of $\widetilde{Q}_b$ and $\widetilde{\zeta}_b$, see equations (\ref{DefnEqn-Rb}) and (\ref{DefnEqn-A}) respectively. The value of $a>0$ is assumed sufficiently small for the proof of Lemma \ref{Lemma-Lowerbound}, relative to some universal constant. The earlier proof of Lemma \ref{Lemma-LyapounovFunc} is conditioned on the eventual choice of $\eta < \frac{a}{C_0}$, for another universal constant $C_0 > 0$, see equation (\ref{Proof-LyapounovFunc-eqn2}). These choices affect the class of initial data ${\mathcal P}$ both by setting the profiles $\widetilde{Q}_b$ and by forcing an upper bound on the value of $\alpha^*$.
\item $\sigma_1$, $\sigma_2$ and $\sigma_3$:
Parameters in the statements of Lemma \ref{Lemma-lambdaIntegralCrude}, Lemma \ref{Lemma-lambdaIntegral}, and Corollary \ref{Corollary-lambdaIntegral2version2}. Their value is chosen (repeatedly) according to circumstance.
\item $\sigma_4$: Defined for Lemma \ref{Lemma-NSobolev}. Value is uniform over all $m>0$ sufficiently small.
\item $\delta_5$: A fixed arbitrary universal constant $0<\delta_5\ll 1$, used in the proof of Lemma \ref{Lemma-NSobolev}.
\item $m'$: Existence of $m'<m$ with particular properties in a key assertion of Proposition \ref{Prop-Improv}, below. Some particular value $m'\in(m-\frac{\sigma_4}{2},m)$ is chosen for the proof of Lemma \ref{Lemma-ControlledHnk}.
\item $\sigma_5$: Parameter in the statement of Lemma \ref{Lemma-NdimEndpointSobolev}. Value is fixed for the proof of Lemma \ref{Lemma-Annular12}.
\item $\sigma_6$: Defined for Lemma \ref{Lemma-AnnularN2}. Value is uniform over all $m>0$ sufficiently small.
\item $m$: Fixed constant $m>0$ features in the bootstrap hypotheses of Definition \ref{Defn-Hypo}, below. For the purpose of various proofs in Chapter \ref{Section-BootAtInfty}, $m$ will be assumed sufficiently small.
The exact value of $m$ may be determined apriori, and will affect the class of initial data ${\mathcal P}$ by forcing an upper bound on the value of $\alpha^*$.
\item $\alpha^*$: Fixed constant $\alpha^*>0$ is determined last. For the purpose of various proofs throughout this paper, $\alpha^*$ will be assumed sufficiently small.
\end{itemize}
\end{remark}
The following bootstrap hypotheses are possible due to our choice of data in ${\mathcal P}$.
\begin{definition}[Time $T_{hyp} > 0$ \& Bootstrap Hypotheses]\label{Defn-Hypo}
Let $0 < T_{hyp} \leq T_{max}$ be the maximum time such that for all $t\in[0,T_{hyp})$ the following two sets of conditions hold:
\begin{description}
\item[Singularity remains of a log-log nature:]
\item[H1.1] {\it Profile remains focused near a singular ring,}
\begin{equation}\label{Hypo1-rz}
\abs{(r(t),z(t))-(1,0)} < \left(\alpha^*\right)^\frac{1}{2}.
\end{equation}
\item[H1.2] {\it Profile remains close to Q near the singular ring,}
\begin{equation}\label{Hypo1-b}\begin{aligned}
0 < b(t) + \norm{\tilde{u}(t)}_{L^2(\real^{3})} < \left(\alpha^*\right)^\frac{1}{10}.
\end{aligned}\end{equation}
\begin{equation}\label{Hypo1-epsilon}\begin{aligned}
\int{\abs{\grad_y\epsilon(t)}^2\mu_{\lambda(t),r(t)}(y)\,dy} + \int_{\abs{y}\leq\frac{10}{b(t)}}{\abs{\epsilon(t)}^2e^{-\abs{y}}\,dy} \leq \Gamma_{b(t)}^\frac{3}{4}.
\end{aligned}\end{equation}
\item[H1.3] {\it Conformal and scaling parameters remain consistent with log-log blowup speed,}
\begin{equation}\label{Hypo1-loglog}\begin{aligned}
\frac{\pi}{10}\frac{1}{\log s} < b(s) < \frac{10\pi}{\log s}, &&
e^{-e^\frac{10\pi}{b(s)}} < \lambda(s) < e^{-e^{\frac{\pi}{10}\frac{1}{b(s)}}}.
\end{aligned}\end{equation}
\item[H1.4] {\it Energy and localized momentum remain normalized,}
\begin{equation}\label{Hypo1-enerMoment}
\lambda^2(t)\abs{E_0} + \lambda(t)\abs{Im\left(\int{\grad\psi^{(x)}\cdot\grad u(t)\overline{u}(t)}\right)}
< \Gamma_{b(t)}^2.
\end{equation}
\item[H1.5] {\it Norm growths are almost monotonic,}
\begin{equation}\label{Hypo1-almostMonotony}\begin{aligned}
\forall s_a \leq s_b \in[s_0,s_1], && \lambda(s_b) \leq 3\lambda(s_a).
\end{aligned}\end{equation}
\item[Regularity away from the singularity persists:]
\item[H2.1] {\it Growth of $\dot{H}^3$ is near scaling,}
\begin{equation}\label{Hypo2-Hnk}
\norm{u(t)}_{H^{3}(\real^{3})} < \frac{e^{+\frac{m}{b(t)}}}{\lambda^{3}(t)},
\end{equation}
\item[H2.2] {\it Strong hierarchy of regularity away from the singular ring persists,}
\begin{equation}\label{Hypo2-Hlower}
\norm{\chi u(t)}_{H^{3-\kappa}} < \frac{e^{+(1+\kappa)\frac{m}{b(t)}}}{\lambda^{3-2\kappa}(t)},
\end{equation}
for each half integer $\frac{1}{2} \leq \kappa < \frac{3}{2}$,
\begin{equation}\label{Hypo2-HlowerN2}
\norm{\chi u(t)}_{H^\frac{3}{2}} < e^{+\frac{2m + \pi}{b(t)}},
\end{equation}
and,
\item[H2.3] {\it Lower-order norms away from the singular ring remain bounded,}
\begin{equation}\label{Hypo2-Hcrit}
\norm{\chi u(t)}_{H^1} < \left(\alpha^*\right)^\frac{1}{10}.
\end{equation}
\end{description}
\end{definition}
\noindent An important consequence of {\bf H1.2}, {\bf H1.3},
and the forthcoming estimate on $\Gamma_b$, (\ref{Eqn-GammaBEstimate}), is that,
\begin{equation}\label{Hypo1-consequenceForLambdaGamma}
\lambda(t) < e^{-e^\frac{\pi}{10b(t)}} < \Gamma_{b(t)}^{10}.
\end{equation}
Therefore as a consequence of {\bf H1.1}
and the forthcoming definition of $A$, (\ref{DefnEqn-A}),
\begin{equation}\label{Hypo1-consequenceForMu}\begin{aligned}
\frac{2}{3} \leq \mu(y) \leq \frac{3}{2}
&&\forall \abs{y}\leq 5 A(t).
\end{aligned}\end{equation}
The region $\abs{y}\leq 5A(t)$ is exceptionally wide, encompassing the support of both the central profile $\widetilde{Q}_b$ and the associated radiation $\widetilde{\zeta}_b$.
\begin{remark}[Geometric decomposition is well defined]
Hypotheses {\bf H1} easily satisfy the conditions of Lemma \ref{Lemma-GeoDecompFixed}, ensuring that $T_{hyp}\leq T_{geo}$ and the unique geometric decomposition (\ref{Eqn-GeoDecomp}) is available.
\end{remark}
\begin{remark}[Higher dimensions]\label{Remark-HigherDim-Hypo}
For the higher dimensional case, extend the hypotheses {\bf H2} to include:
\begin{itemize}
\item Controlled growth of $\norm{u(t)}_{H^{N}}$, in place of {\bf H2.1}
\item Equation (\ref{Hypo2-Hlower}) for each half integer $\frac{1}{2} \leq \kappa < \frac{N}{2}$, and phrase (\ref{Hypo2-HlowerN2}) with respect to $H^\frac{N}{2}$.
\item Bounded behaviour in $H^{\frac{N-1}{2}}$ away from the singular ring, in place of {\bf H2.3}.
\end{itemize}
The obvious changes to the class of initial data should also be made.
\end{remark}
\begin{comment}
\begin{remark}
Equation (\ref{Hypo1-loglog}) encodes to what degree it is that $b_s \approx -e^{-\frac{C}{b}}$ and $\frac{\lambda_s}{\lambda} \approx -b$, respectively. That is, to what degree $u(t)$ obeys the log-log blowup rate.
The almost-monotonous property of equation (\ref{Hypo1-almostMonotony}) simply demands that we are in the 'final sequence' of blowup. Equation (\ref{Hypo1-almostMonotony}) is equivalent to strict monotony of the time-average of parameter $\lambda$ over appropriate slices.
\end{remark}
\begin{remark}
The first key outcome of Merle-Rapha\"el log-log analysis will be an (improved) bound of the type (\ref{Hypo1-epsilon}). Such a bound ensures that any failure of local wellposedness must be described by the dynamics of the parameters.
\end{remark}
\end{comment}
\begin{prop}[Bootstrap Conclusion]\label{Prop-Improv}
For $\alpha^*>0$ sufficiently small, hypotheses (\ref{Hypo1-rz}) through (\ref{Hypo2-Hcrit}) are not sharp. There exists $m'<m$ such that for all $t\in[0,T_{hyp})$:
\begin{description}
\item[I1.1]
\begin{equation}\label{Improv1-rz}
\abs{(r(t),z(t))-(1,0)} < \left(\alpha^*\right)^\frac{2}{3},
\end{equation}
\item[I1.2]
\begin{equation}\label{Improv1-b}\begin{aligned}
0 < b(t) + \norm{\tilde{u}(t)}_{L^2(\real^{3})} < \left(\alpha^*\right)^\frac{1}{5},
\end{aligned}\end{equation}
\begin{equation}\label{Improv1-epsilon}\begin{aligned}
\int{\abs{\grad_y\epsilon(t)}^2\mu_{\lambda(t),r(t)}(y)\,dy} + \int_{\abs{y}\leq\frac{10}{b(t)}}{\abs{\epsilon(t)}^2e^{-\abs{y}}\,dy} \leq \Gamma_{b(t)}^\frac{4}{5},
\end{aligned}\end{equation}
\item[I1.3]
\begin{equation}\label{Improv1-loglog}\begin{aligned}
\frac{\pi}{5}\frac{1}{\log s} < b(s) < \frac{5\pi}{\log s}, &&
e^{-e^\frac{5\pi}{b(t)}} < \lambda(t) < e^{-e^{\frac{\pi}{5}\frac{1}{b(t)}}},
\end{aligned}\end{equation}
\item[I1.4]
\begin{equation}\label{Improv1-enerMoment}
\lambda^2(t)\abs{E_0} + \lambda(t)\abs{Im\left(\int{\grad\psi^{(x)}\cdot\grad u(t)\overline{u}(t)}\right)}
< \Gamma_{b(t)}^4,
\end{equation}
\item[I1.5]
\begin{equation}\label{Improv1-almostMonotony}\begin{aligned}
\forall s_a \leq s_b \in[s_0,s_1], && \lambda(s_b) \leq 2\lambda(s_a).
\end{aligned}\end{equation}
\item[I2.1]
\begin{equation}\label{Improv2-Hnk}
\norm{u(t)}_{H^{3}(\real^{3})} < \frac{e^{+\frac{m'}{b(t)}}}{\lambda^{3}(t)},
\end{equation}
\item[I2.2]
\begin{equation}\label{Improv2-Hlower}
\norm{\chi u(t)}_{H^{3-\kappa}} < \frac{e^{+(1+\kappa)\frac{m'}{b(t)}}}{\lambda^{3-2\kappa}(t)},
\end{equation}
for each half integer $\frac{1}{2} \leq \kappa < \frac{3}{2}$,
\begin{equation}\label{Improv2-HlowerN2}
\norm{\chi u(t)}_{H^\frac{3}{2}} < e^{+\frac{2m' + \pi}{b(t)}},
\end{equation}
and,
\item[I2.3]
\begin{equation}\label{Improv2-Hcrit}
\norm{\chi u(t)}_{H^1} < \left(\alpha^*\right)^\frac{1}{5}.
\end{equation}
\begin{comment}
[Consequences of Proposition \ref{Prop-Improv2}]\label{Remark-Prop2WillShowT2=T1}
From Proposition \ref{Prop-Improv1}, we already know that the geometric decomposition will continue so long as (\ref{Hypo2-Hcrit}) holds. Thus $T_2 \leq {\mathcal T}_1$ and Proposition \ref{Prop-Improv2} proves that $t_2 = T_2$, which implies from local wellposedness and Definition \ref{Defn-Hypo1} that either $b(t) \to 0$ or $\lambda(t) \to 0$ as $t\rightarrow T_2$, forcing us to conclude that $t_1 \leq T_2$ and $T_1 = T_2$.
See Chapter \ref{Section-FinalProof}.
\end{comment}
\end{description}
As a consequence, $T_{hyp} = T_{max}$.
\end{prop}
\subsection{Strategy of Proof: the log-log argument}
\label{Subsec-StrategyPart1}
\begin{comment}
The groundbreaking series of work \cite{MR-BlowupDynamic-05}, \cite{MR-SharpUpperL2Critical-03}, \cite{MR-UniversalityBlowupL2Critical-04} and \cite{MR-SharpLowerL2Critical-06} are the basis for describing the behaviour of the singular part of the solution. Suppose that we may decompose a solution $v(t)$ of the two-dimensional cubic NLS as,
\begin{equation}\label{Strategy1-decomp}
v(t,r,z) = \frac{1}{\lambda(t)}\left(Q+\epsilon_v\right)\left(\frac{(r(t),z(t))-(r,z)}{\lambda}\right)e^{-i\gamma(t)},
\end{equation}
for some (continuous) parameters $\lambda(t)$, $r(t)$, $z(t)$, $\gamma(t)$. Then, we seek to reduce the question of blowup dynamics to finite-dimensional ODE dynamics for the parameters of decomposition (\ref{Strategy1-decomp}). This is only possible due to the algebraic structure associated with $Q$. We will denote $\Lambda$ the operator,
\begin{equation}\label{DefnEqn-Lambda}\begin{aligned}
\Lambda = 1+y\cdot\grad,
&& \text{ which satisfies, }
&& \left(f,\Lambda g\right) = -\left(\Lambda f,g\right).
\end{aligned}\end{equation}
One might recognize $\Lambda$ from either the argument $E(Q) = 0$:
\begin{equation}\label{Strategy1-energy}
\left(0,\Lambda(Q)\right) = \left(\laplacian Q-Q+Q\abs{Q}^2,\Lambda(Q)\right) = -2E(Q),
\end{equation}
or from the Pohozaev identity:
\begin{equation}\label{Strategy1-pohozaev}
\left(0,\Lambda(v)\right)
= Re\left(iv_s+\laplacian_y v + v\abs{v}^2,\Lambda(v)\right)
= \frac{1}{2}\frac{d}{ds}Im\int{v\,y\cdot\grad\overline{v}\,dy} - 2E(v),
\end{equation}
which is also a consequence of formally calculating the virial identity, $\frac{d^2}{d^2s}\int{\abs{y}^2\abs{v}^2\,dy}$. The linearized propagator near $Q$ is a matrix operator on real and imaginary parts,
\begin{equation}\label{DefnEqn-L}\begin{aligned}
L(\epsilon) =
\left\lbrack\begin{smallmatrix}L_+&0\\0&L_-\end{smallmatrix}\right\rbrack
\left\lbrack\begin{smallmatrix}\epsRe\\\epsIm\end{smallmatrix}\right\rbrack
&& \text{ with: }
&& \begin{aligned}
L_+ &= -\laplacian + 1 - 3Q^2\\
L_- &= -\laplacian + 1 - Q^2
\end{aligned}
\end{aligned}
\end{equation}
That $e^{itL}$ has exactly six growing modes was established \cite{Weinstein85} - a precursor of the Spectral Property below. Weinstein also noted the following essential algebraic property:
\begin{equation}\label{Strategy1-Algebra}
L_+\left(\Lambda Q\right) = -2Q.
\end{equation}
Substitute ansatz (\ref{Strategy1-decomp}) into cubic NLS, introduce rescaled variables in time, $s(t)=\int^t{\frac{1}{\lambda^2(\tau)}\,d\tau},$ and in space, $y(t) = \frac{(r,z) -\left(r(t),z(t)\right)}{\lambda}$. With new notation $\tilde{\gamma} = -s - \gamma$, the equation for $\epsilon$ is,
\begin{equation}\label{Strategy1-epsEqn}
i\epsilon_s -L(Q+\epsilon)+\left(\epsilon^2Q+2\abs{\epsilon}^2Q\right)+\epsilon\abs{\epsilon}^2
=
i\left(\frac{\lambda_s}{\lambda}y + \frac{(r_s,z_s)}{\lambda}\right)\cdot\grad(Q+\epsilon)
+\tilde{\gamma}_s\left(Q+\epsilon\right).
\end{equation}
The RHS of (\ref{Strategy1-epsEqn}) involves exactly the type of dynamics we wish to focus upon; to manage the LHS of (\ref{Strategy1-epsEqn}) we revisit the Pohozaev calculation (\ref{Strategy1-pohozaev}). In anticipation of an inner product of (\ref{Strategy1-epsEqn}) with $\Lambda Q$, note the following crucial use of algebraic property (\ref{Strategy1-Algebra}):
\begin{equation}\label{Strategy1-epsLinear}\begin{aligned}
Re\left(-L(Q+\epsilon),\Lambda Q\right) = -E(Q) +& Re\left(-L(\epsilon),\Lambda Q\right) = -E(Q)+2Re\left(\epsilon,Q\right).
\end{aligned}
\end{equation}
The final term of (\ref{Strategy1-epsLinear}) also appears when we re-write the conservation of energy (\ref{ConserveEnergy}) in our rescaled variables,
\begin{equation}\label{Strategy1-epsEnergy}\begin{aligned}
2\lambda^2E_0
&= \int{\abs{\grad_y(Q+\epsilon)}^2} - \frac{1}{2}\int{\abs{Q+\epsilon}^4}\\
&= 2E(Q) - 2Re\left(\epsilon,Q\right) + \int{\abs{\grad_y\epsilon}^2}
-\left(
Re\int{\epsilon^2Q^2}+2\int{\abs{\epsilon}^2Q^2}+O(\epsilon^3)
\right)
\end{aligned}\end{equation}
where the second line followed by parts and substitution of (\ref{DefnEqn-Q}). An estimate for the terms $O(\epsilon^3)$ is the main difficulty in extending the two-dimensional log-log argument of \cite{MR-SharpUpperL2Critical-03}, \cite{MR-SharpLowerL2Critical-06} to the higher-dimensional setting of Theorem \ref{Thm-MainResult}.
Still in anticipation of an inner product with $\Lambda Q$, note that by parts,
\begin{equation}\label{Strategy1-epsQuad}
Re\left(\epsilon^2Q+2\abs{\epsilon}^2Q ,\Lambda Q\right)
= \frac{1}{2}Re\left(\epsilon^2+2\abs{\epsilon}^2,\Lambda (Q^2)\right)
\end{equation}
To the order of $\epsilon^2$, the sum of (\ref{Strategy1-epsEnergy}) with (\ref{Strategy1-epsQuad}) is the Hamiltonian linearized about $Q$:
\begin{equation}\label{DefnEqn-H}
\begin{aligned}
H(\epsilon,\epsilon) =
\left\lbrack\begin{smallmatrix}{\mathcal L}_\text{re}&0\\0&{\mathcal L}_\text{im}\end{smallmatrix}\right\rbrack
\left\lbrack\begin{smallmatrix}\epsRe\\i \epsIm\end{smallmatrix}\right\rbrack
\cdot
\left\lbrack\begin{smallmatrix}\epsRe\\-i \epsIm\end{smallmatrix}\right\rbrack
&& \text{ with: }
&& \begin{aligned}
{\mathcal L}_\text{re} &= -\laplacian + 3Qy\cdot\grad Q\\
{\mathcal L}_\text{im} &= -\laplacian + Qy\cdot\grad Q
\end{aligned}
\end{aligned}
\end{equation}
The link between $\Lambda Q$ and the linearized Hamiltonian should be no surprise, especially after (\ref{Strategy1-energy}). It is more surprising that $H$ enjoys stable modes much as Weinstein found for $L$:
\begin{prop}[Spectral Property]
\label{Prop-SpectralProperty}
There exists a universal constant ${\delta}_0 > 0$ such that $\forall \epsilon \in H^1$:
\begin{equation}\label{Eqn-2DSpectral}\begin{aligned}
H(\epsilon,\epsilon) \geq &{\delta}_0
\left(\int_{y\in\real^2}{\abs{\grad_y\epsilon}^2} +
\int_{y\in\real^2}{\abs{\epsilon^2}e^{-\abs{y}}}\right)\\
&-\frac{1}{{\delta}_0}\left(
\begin{aligned}&Re\left(\epsilon,Q\right) + Re\left(\epsilon,\Lambda Q\right) +
Re\left(\epsilon,yQ\right)\\ &+ Im\left(\epsilon,\Lambda Q\right) +
Im\left(\epsilon,\Lambda^2Q\right) + Im\left(\epsilon,\grad Q\right)\end{aligned}
\right)^2.
\end{aligned}\end{equation}
\end{prop}
The Spectral Property for one-dimensional ground state is proven \cite[Proposition 2]{MR-BlowupDynamic-05}. The two-dimensional version stated here has a numerical proof \cite{FMR-ProofOfSpectralProperty-06}. Assume, momentarily, that by some means we can ensure $H$ is coercive. Then, summing (\ref{Strategy1-epsLinear}), (\ref{Strategy1-epsEnergy}) and (\ref{Strategy1-epsQuad}), the real part of an inner product of (\ref{Strategy1-epsEqn}) with $\Lambda Q$ produces:
\begin{equation}\label{Strategy1-localVirial}
\left(\int_{y\in\real^2}{\abs{\grad_y\epsilon}^2} + \int_{y\in\real^2}{\abs{\epsilon^2}e^{-\abs{y}}}\right) \lesssim
\partial_sIm\left(\epsilon,\Lambda Q\right)
+ 2\lambda^2E_0
+ O(\epsilon^3)
+ O(\text{parameters}),
\end{equation}
which we style a `virial estimate' by virtue of inspiration (\ref{Strategy1-pohozaev}). Regarding the negative modes of $H$:
\begin{itemize}
\item Parameters of the decomposition (\ref{Strategy1-decomp}) may be modulated to ensure
$Re\left(\epsilon,\Lambda Q\right)$, $Re\left(\epsilon,yQ\right)$, and $Im\left(\epsilon,\Lambda^2Q\right)$ are zero.
\item Conservation of energy (\ref{ConserveEnergy}) will control $Re\left(\epsilon,Q\right)$ by $O(\epsilon^2)$, which affects (\ref{Strategy1-localVirial}) as a lower-order term $O(\epsilon^4)$.
\item Conservation of momentum (\ref{ConserveMoment}) will likewise control $Im\left(\epsilon,\grad Q\right)$.
\end{itemize}
The method thus far exposed is essentially that of \cite{MR-BlowupDynamic-05} - it does not give the sharp log-log upper bound, as we have insufficient control on $Im\left(\epsilon,\Lambda Q\right)$. Motivated by the heuristics of \cite{SulemSulem}, \cite{MR-SharpUpperL2Critical-03} introduces a one-parameter family of almost self-similar profiles close to $Q$ both structurally and in $H^1$, and indexed by $b$. These new profiles, in place of $Q$, lend an extra parameter to our geometric decomposition with which we may enforce $Im\left(\epsilon,\Lambda Q\right)=0$.
The evolution of this new parameter, $\partial_sb$, will appear in place of $\partial_sIm\left(\epsilon,\Lambda Q\right)$ in equation (\ref{Strategy1-localVirial}), along with various terms due to the structure of the almost self-similar profiles. These structural properties are central to proving the log-log blowup rate.
While our estimate analogous to (\ref{Strategy1-localVirial}) will give sufficient information to establish the sharp log-log upper bound for the blowup rate \cite{MR-SharpUpperL2Critical-03}, we will nevertheless only have good $H^1$ control for $\epsilon$ at those times where $\partial_s b \leq 0$. To bridge the periods between these times, we follow \cite{MR-SharpLowerL2Critical-06} to establish a (decreasing) Lyapounov functional ${\mathcal J}$ that satisfies two astonishing estimates:
\begin{enumerate}
\item To leading order (using conservation of mass), ${\mathcal J}$ is the excess, unejected mass that remains in the focusing regime.
\item Strictly accounting for the mass of the almost self-similar profiles (and using conservation of energy), ${\mathcal J}$ is the size of $\epsilon$ in $H^1$.
\end{enumerate}
For the reader recently acquainted with the techniques of
\cite{MR-BlowupDynamic-05}, \cite{MR-SharpUpperL2Critical-03}, \cite{MR-UniversalityBlowupL2Critical-04} and \cite{MR-SharpLowerL2Critical-06}, the following technicalities will be fully addressed in Chapter \ref{Section-BootLoglog}:
\begin{enumerate}
\item {\it Advances in the Merle-Rapha\"el Argument}
\begin{enumerate}
\item {\it Different Orthogonality Conditions}
We will make a geometric decomposition that demands $Re\left(\epsilon,\abs{y}^2Q\right) = 0$, instead of $Re\left(\epsilon,\Lambda Q\right)=0$. This alternate choice of orthogonality condition, introduced in \cite{MR-UniversalityBlowupL2Critical-04}, allows a tighter estimate of $\gamma(t)$. See Lemma \ref{Lemma-AlternateSpectral}.
\item {\it Linearized Operator about $\widetilde{Q}_b$}
We will introduce a new evolution operator linearized about the almost self-similar profiles, equation (\ref{DefnEqn-M}). We avoid proving a new Spectral Property by estimating the difference between the new linearized Hamiltonian and $H$. See equation (\ref{DefnEqn-HbTilde}).
\item {\it Repeated Virial Argument}
To begin, we will prove a virial identity built around the almost self-similar profiles, analogous to (\ref{Strategy1-localVirial}) - see Lemma \ref{Lemma-LocalVirial}. We will then essentially repeat the argument, proving a virial identity built around both the self-similar profiles and an approximation of their radiative tails - see Lemma \ref{Lemma-RefinedVirial}. This second virial estimate does not in itself offer better control of $\epsilon$ than did Lemma \ref{Lemma-LocalVirial}; rather we will trade terms of the `bad sign' with a mass-dispersion calculation to derive the Lyapounov functional.
\end{enumerate}
\item {\it Effects of Cylindical Symmetry}
Our rescaled variable $y$ is two-dimensional, while our problem is posed in $\real^{3}$. This induces a cylindrical measure of integration $\mu(y) = 2\pi r = 2\pi \left(r(t)+\lambda(t) y_1\right)$, which
enters into all the integrals of the conservation of energy (\ref{Strategy1-epsEnergy}). We will routinely split the measure as $r(t) \sim 1$, for which we follow the two-dimensional argument outlined above,
and a variable part that will be negligible on the support of the singular profile. See (\ref{Hypo1-consequenceForLambdaGamma}) below.
\item {\it Global Terms in $\epsilon$}
Terms of order $\epsilon^3$ involve an interaction with $Q$ that will localize them to a region near the singular ring $(r,z)\sim (1,0)$. Then $\mu(y) \sim 1$ and we may apply two-dimensional Sobolev estimates. However, the term $\int_{\real^{2}}{\abs{\epsilon(y)}^4\mu(y)}$ arising from the conservation of energy (\ref{Strategy1-epsEnergy})
is truly global and not of a two-dimensional nature.
To prove log-log behaviour, we will assume the estimate,
\[
\int_{\real^{2}}{\abs{\epsilon(y)}^4\mu(y)} << \int_{\real^2}{\abs{\grad_y\epsilon}^2\mu(y)},
\] for the duration of Chapter \ref{Section-BootLoglog}. The goal of Chapter \ref{Section-BootAtInfty} is then to establishing this estimate through persistence of regularity.
\end{enumerate}
\end{comment}
We will establish statements {\bf I1} in Chapter \ref{Section-BootLoglog} using the arguments of \cite{MR-SharpUpperL2Critical-03} and \cite{MR-SharpLowerL2Critical-06}. Here we identify the main challenge in maintaining the log-log dynamics.
As with all modulation arguments, we seek to reduce the question of blowup to a finite-dimensional ODE dynamic for the parameters. This is only possible due to the algebraic structure associated with $Q$.
\begin{comment}
We will denote $\Lambda$ the operator,
\begin{equation}\label{DefnEqn-Lambda}\begin{aligned}
\Lambda = 1+y\cdot\grad,
&& \text{ which satisfies, }
&& \left(f,\Lambda g\right) = -\left(\Lambda f,g\right).
\end{aligned}\end{equation}
\end{comment}
Recall the operator $\Lambda = 1 +y\cdot\grad_y$, which one might recognize from either the argument $E(Q) = 0$:
\begin{equation}\label{Strategy1-energy}
\left(0,\Lambda(Q)\right) = \left(\laplacian Q-Q+Q\abs{Q}^2,\Lambda(Q)\right) = -2E(Q),
\end{equation}
or from the Pohozaev identity:
\begin{equation}\label{Strategy1-pohozaev}
\left(0,\Lambda(v)\right)
= Re\left(iv_s+\laplacian_y v + v\abs{v}^2,\Lambda(v)\right)
= \frac{1}{2}\frac{d}{ds}Im\int{v\,y\cdot\grad\overline{v}\,dy} - 2E(v),
\end{equation}
which is also a consequence of formally calculating the virial identity, $\frac{d^2}{d^2s}\int{\abs{y}^2\abs{v}^2\,dy}$.
Substitution of (\ref{Eqn-GeoDecomp}) into (\ref{Eqn-NLS}) will produce an equation for $\epsilon$. Ignoring the distinction between $Q$ and $\Qb$, the terms linear in $\epsilon$ are, $i\partial_s\epsilon + L(\epsilon)$, where $L$ is the linearized propagator near $Q$. As a matrix on real and imaginary parts,
\begin{equation}\label{DefnEqn-L}\begin{aligned}
L(\epsilon) =
\left.
\begin{bmatrix}0&L_-\\-L_+&0\end{bmatrix}
\begin{bmatrix}\epsRe\\i\, \epsIm\end{bmatrix}
\right.
&& \text{ with: }
&& \begin{aligned}
L_+ &= -\laplacian + 1 - 3Q^2\\
L_- &= -\laplacian + 1 - Q^2
\end{aligned}
\end{aligned}
\end{equation}
Weinstein noted \cite{Weinstein85} the following:
\begin{equation}\label{Strategy1-Algebra}\begin{aligned}
L_-\left(\abs{y}^2Q\right) = -2\Lambda Q, &&
L_-\left(yQ\right) = -2\grad Q, && \text{ and } &&
L_+\left(\Lambda Q\right) = -2Q.
\end{aligned}\end{equation}
These algebraic properties are the inspiration of the orthogonality conditions so that, by taking appropriate inner products of the $\epsilon$-equation, linear terms cancel. For example, the imaginary part of the inner product with $\abs{y}^2Q$ has no linear terms due to conditions (\ref{Eqn-GeoDecomp-OrthogReY2}) and (\ref{Eqn-GeoDecomp-OrthogImL}). The imaginary part of the inner product with $yQ$ is controlled by momentum.
The most fruitful calculation is when we take the real part of an inner product of the $\epsilon$-equation with $\Lambda Q$. This is of course a localized version of equation (\ref{Strategy1-pohozaev}). We substitute conservation of energy to eliminate the linear term, $2\,Re\left(\epsilon,Q\right)$, which is due to the third identity of equation (\ref{Strategy1-Algebra}). The remaining terms quadratic in $\epsilon$ form the following,
\begin{equation}\label{DefnEqn-H}
\begin{aligned}
H(\epsilon,\epsilon) =
\left.
\begin{bmatrix}{\mathcal L}_\text{re}&0\\0&{\mathcal L}_\text{im}\end{bmatrix}
\begin{bmatrix}\epsRe\\i\, \epsIm\end{bmatrix}
\cdot
\begin{bmatrix}\epsRe\\-i\, \epsIm\end{bmatrix}
\right.
&& \text{ with: }
&& \begin{aligned}
{\mathcal L}_\text{re} &= -\laplacian + 3Qy\cdot\grad Q\\
{\mathcal L}_\text{im} &= -\laplacian + Qy\cdot\grad Q
\end{aligned}
\end{aligned}
\end{equation}
Operator $H(\epsilon,\epsilon)$ is the derivative with respect to scaling of the conserved energy of the linear flow. It has coercivity properties that mirror the stability of $Q$:
\begin{prop}[Spectral Property]
\label{Prop-SpectralProperty}
There exists a universal constant ${\delta}_0 > 0$ such that $\forall v \in H^1$:
\begin{equation}\label{Eqn-2DSpectral}\begin{aligned}
H(v,v) \geq &{\delta}_0
\left(\int_{y\in\real^2}{\abs{\grad_yv}^2} +
\int_{y\in\real^2}{\abs{v^2}e^{-\abs{y}}}\right)\\
&-\frac{1}{{\delta}_0}\left(
\begin{aligned}&Re\left(v,Q\right) + Re\left(v,\Lambda Q\right) +
Re\left(v,yQ\right)\\ &+ Im\left(v,\Lambda Q\right) +
Im\left(v,\Lambda^2Q\right) + Im\left(v,\grad Q\right)\end{aligned}
\right)^2.
\end{aligned}\end{equation}
\end{prop}
The two-dimensional Spectral Property as stated here has a numerical proof \cite{FMR-ProofOfSpectralProperty-06}\footnotemark.
\footnotetext{
The numerical proof is given for the $L^2$-critical problem in dimensions $N=2,3,4$ and nonlinearity $u\abs{u}^\frac{4}{N}$. Proof for dimension $N=1$ is explicit, \cite[Proposition 2]{MR-BlowupDynamic-05}.
}
Assuming we can ensure $H$ is coercive, the goal is to prove the {\it local virial identity},
\begin{equation}\label{Strategy1-localVirial}
b_s \geq \delta_1\left(\epsNorm\right) - \Gamma_b^{1-C\eta}.
\end{equation}
To prove (\ref{Strategy1-localVirial}) using the Spectral Property requires we control the contribution from all other terms of the conservation of energy. In particular, we must establish the {\bf non-local} control,
\begin{equation}\label{Eqn-HRTemp1}
\int_{\real^{2}}{\abs{\epsilon(y)}^4\mu(y)} \ll \int_{\real^2}{\abs{\grad_y\epsilon}^2\mu(y)}.
\end{equation}
This is our main challenge.
The local virial identity (\ref{Strategy1-localVirial}) is a satisfactory control for $\epsilon$ at times where $b_s < 0$. However, our argument is based on approximating the central profile of the solution, therefore we cannot expect monotinicity in our modulation parameters. Including the radiation $\zb$ to better approximate the central profile, repeating the local virial calculation, and taking into account the mass flux leaving the support of the radiation, Merle \& Rapha\"el discovered a Lyapounov functional, \cite{MR-SharpLowerL2Critical-06}.
It is remarkable that we may approximate the Lyapounov functional very precisely in terms of a positive multiple of a norm of $\epsilon$. The functional is then used to bridge the control of $\epsilon$ between times where $b_s<0$. The approximation here is achieved through the conservation of energy, and involves equation (\ref{Eqn-HRTemp1}) a second time.
Regarding (\ref{Eqn-HRTemp1}), change variables,
\begin{equation}\label{Eqn-HRTemp2}
\int_{\real^{2}}{\abs{\epsilon(y)}^4\mu(y)}
= \lambda^2\int_{\real^{3}}{\abs{\widetilde{u}}^4}
= \lambda^2\int_{\real^{3}}{\abs{\chi\widetilde{u}}^4}
+\lambda^2\int_{\real^{3}}{(1-\chi^4)\abs{\widetilde{u}}^4}.
\end{equation}
Since the support of $\chi$ includes the origin, we must apply three-dimensional Sobolev to that term,
\[
\norm{\chi \widetilde{u}}_{L^4(\real^3)}^4 \lesssim \norm{\chi u}_{H^\frac{1}{2}(\real^3)}^2\norm{\chi u}_{H^1(\real^3)}^2.
\]
Changing variables again, we observe that to achieve (\ref{Eqn-HRTemp1}) requires at least that, $\norm{\chi u}_{H^\frac{1}{2}(\real^3)} \ll 1$.
\begin{remark}[Higher dimensions]\label{Remark-HigherDim-StrategyUseHCrit}
In general the Sobolev embedding into $L^4$ involves the critical norm $H^{\frac{N-2}{2}}$,
\[
\norm{\chi u}_{L^4(\real^N)}^4 \lesssim \norm{\chi u}_{H^\frac{N-2}{2}(\real^N)}^2\norm{\chi u}_{H^1(\real^N)}^2.
\]
\end{remark}
\begin{comment}
\subsection{Strategy of Proof: the use of higher regularity}
\label{Subsec-needRegularity}
In this section we aim to illustrate the utility of higher regularity in bootstrapping a control of the form $\norm{\chi u(t)}_{H^\frac{1}{2}(\real^3)} << 1.$
Let $\varphi$ be any cutoff function supported away from the singular ring. The situation where $\varphi = 1$ on the support of $\grad \chi$, and the situation where $\varphi = \chi$ are of particular interest.
By multiplying NLS, the equation for $\varphi u$ is,
\begin{equation}\label{Eqn-needRegularity-varphiEqn}
i\left(\varphi u\right)_t + \laplacian\left(\varphi u\right)
+ \varphi u \abs{\varphi u}^2
= \left(\varphi^2-1\right)\varphi u\abs{u}^2
+ 2\grad\varphi\cdot\grad u + u\laplacian\varphi.
\end{equation}
Each term presents its own challenge. The most difficult, and the focus for this section, is $\left(\varphi^2-1\right)\varphi u\abs{u}^2$, which we will denote as $\varphi u \abs{\psi u}^2$. For either of the situations we have in mind, $\psi$ will be supported far from $r=0$. Consider several strategies to produce a bound for $\norm{\varphi u}_{H^\nu}$:
\begin{itemize}
\item {\it Direct Calculation}
Substitute (\ref{Eqn-needRegularity-varphiEqn}) into $\frac{d}{dt}\norm{\varphi u}_{H^\nu}^2$ to conclude,
\begin{equation}\label{Remark-ImportanceAnnular12-eqn2}
\frac{d}{dt}\norm{\varphi u}_{H^\nu_x}^2 \lesssim \abs{ Im\left(
\int_x{D^\nu\left(\varphi u \abs{\psi u}^2\right)D^\nu\left(\varphi \overline{u}\right)}\right) }.
\end{equation}
The operator $D^\frac{1}{2}$ does not follow Leibniz properly, so there is no trivial cancellation. The best control that should be hoped for is,
\begin{multline}\label{Eqn-needRegularity-alternSobo}
\abs{\int_x{D^\nu\left(\varphi u \abs{\psi u}^2\right)D^\nu\left(\varphi \overline{u}\right)}}\\
\lesssim \left\{\begin{aligned}
&\norm{\varphi u}_{H^\nu}^2\norm{\psi u}_{L^\infty}^2
&&\text{ or,}\\
&\norm{\varphi u}_{H^\nu}^{2-\frac{\delta}{1-\nu}}\norm{\psi u}_{L^2}^\delta\,
\norm{\varphi u}_{H^1}^{\frac{\delta}{1-\nu}}\norm{\psi u}_{H^1}^{2-\delta}
&&\text{ for any }0<\delta < 2.
\end{aligned}\right.
\end{multline}
Any bootstrap hypothesis we might have for $\varphi u$ cannot help to estimate the behaviour of $u$ on the support of $\varphi^2-1=\psi^2$. In the second case of (\ref{Eqn-needRegularity-alternSobo}) we could achieve,
\[
\norm{\varphi u}_{H^\nu}^\frac{\delta}{1-\nu}
\leq C(\alpha^*) \int^t{\norm{\varphi u}_{H^1}^\frac{\delta}{1-\nu}\norm{\psi u}_{H^1}^{2-\delta}}
\sim \int^t{\frac{1}{\lambda^{2 + \delta\frac{\nu}{1-\nu}}}}
\sim \frac{1}{\lambda^{\delta\frac{\nu}{1-\nu}}}.
\]
(See Lemma \ref{Lemma-lambdaIntegral} to justify the integral.) Note that this bound is exactly the same as interpolation. The first case of (\ref{Eqn-needRegularity-alternSobo}) we are left to conclude,
\begin{equation}\label{Eqn-needRegularity-forcedAnswer}
\norm{\varphi u(t)}_{H^\nu}^2
\lesssim e^{\int^t{\norm{\psi u(\tau)}_{L^\infty_x}^2\,d\tau}}.
\end{equation}
In the radially-symmetric case \cite{R06}, the RH side of (\ref{Eqn-needRegularity-forcedAnswer}) is a constant, through Strauss' radial interpolation and the special time-integrability of the $H^1$ norm away from a log-log singularity - see (\ref{Eqn-epsIntegral}).
\item {\it Strichartz}
Direct calculation gave $\norm{\chi u}_{L^\infty_tH^\nu_x} \lesssim \norm{\chi u\abs{\psi u}^2}_{L^1_tH^\nu_x}$. This is an admissible Strichartz estimate. Assuming such an estimate for different admissible pairs does not provide any improvement, as the numerology of time and spatial norms perfectly cancel out - the power of $\frac{1}{\lambda}$ integrated in time is constant.
\item {\it Kato Smoothing}
With Kato's smoothing effect we would expect,
\begin{equation}\label{Remark-ImportanceAnnular12-eqn5}
\norm{\chi u}_{L^\infty_tH^{\nu}_x} \lesssim \norm{\chi u \abs{\psi u}^2}_{L^2_tH^{\nu-\frac{1}{2}}_x}.
\end{equation}
Ignoring any technicalities, the best control that should be hoped for is,
\[
\norm{\chi u\abs{\psi u}^2}_{H^{\nu-\frac{1}{2}}_x}
\lesssim \norm{\chi u}_{H^\nu}\norm{\psi u}_{H^\frac{3}{4}}^2.
\]
To achieve an estimate, we must hope for $\int^t{\norm{\psi u}_{H^\frac{3}{4}}^4}<<1$, which is unlikely.
\end{itemize}
In this paper, we will introduce and control higher regularity so as to perform a calculation almost identical to (\ref{Eqn-needRegularity-forcedAnswer}) - see Lemma \ref{Lemma-Annular12}.
\end{comment}
\subsection{Strategy of Proof: persistence of regularity}
\label{Subsec-StrategyPart2}
\begin{comment}
Let us consider the $L^2$-critical two-dimensional cubic equation.
It is known that the residual profile $\widetilde{u}(t) \to u^*$ at blowup time exists in $H^s$ for $s\leq 0$ only \cite{MR-ProfilesQuantization-05}. It is our belief that $\widetilde{u}$ loses regularity precisely from the radiation term $\widetilde{\zeta_b}$ that is coupled directly to the focusing singularity. While the support of $\widetilde{\zeta_b}$ is unbounded in $y$ as $t\rightarrow T_{max}$, in the original variables $(r,z)$ the support $\abs{x} \lesssim A(t)\lambda(t)$ collapases to the singular point - see (\ref{DefnEqn-A}).
\begin{remark}[Crude Heuristics]
Let us consider the radiation exclusively in the region far from the coupling with the singular profile - that is, the region where we might expect a coupling between the radiation and the dispersive regime. Here, in the original variables, the radiation has $\dot{H}^1 \approx \frac{\Gamma_b^{\frac{1}{2}-}}{\lambda}$ and mass $L^2 \approx \Gamma_b^{\frac{1}{2}-}$ - see (\ref{Eqn-zb_smallH1}) and (\ref{Eqn-GammaBEstimate}). If we suppose that the radiation only induces dispersion on Littlewood-Paley blocks of frequency $\sim \frac{1}{\lambda(t)}$, then from the log-log blowup rate there is not nearly enough time for those blocks to disperse from the support of the radiation, $\abs{x} \lesssim A(t)\lambda(t)$, into a fixed region, $\abs{x} > C > 0$, before blowup at time $T_{max}$.
\end{remark}
\begin{remark}[Slower now, Faster later]
It is common parlance to say log-log blowups $\sqrt{\frac{\log\abs{\log(T_{max}-t)}}{T_{max}-t}}$ are `slow' compared with $\frac{1}{T_{max}-t}$ blowups. Note however that, given a blowup of each type in their `final sequence' - that is, two solutions initially with the same (large) size in $H^1$ - then {\bf the log-log blowup occurs first}.
\end{remark}
\end{comment}
Once we have established the log-log nature of our blowup, we expect powers of $\frac{1}{\lambda}$ to be as integrable in time as powers of $\sqrt{\frac{\log\abs{\log(T_{max}-t)}}{T_{max}-t}}$. Indeed, as noted in \cite{RaphaelSzeftel-StandingRingNDimQuintic-08},
\begin{equation}\label{Strategy2-lambdaIntegral}
\int_0^t{\frac{d\tau}{\lambda^{2\nu+1}(\tau)}} \leq C(\epsilon)\frac{1}{\lambda^{2\nu-1+\epsilon}(t)},
\end{equation}
for any $\delta > 0$ and $\nu \geq \frac{1}{2}$. As we explain below, our argument requires more care. We prove that,
\begin{equation}\begin{aligned}\label{Strategy2-lambdaIntegralImproved}
\int_0^t{\frac{\abs{\log\lambda}^{\sigma^*}}{\lambda^{2\nu+1}(\tau)}\,d\tau}
\leq C(\sigma^*,\sigma,\nu) \frac{\abs{\log\lambda}^{\sigma}(t)}{\lambda^{\nu-1}(t)},
\end{aligned}\end{equation}
for any $\sigma^*<\sigma$, of either sign, and $\nu \geq \frac{1}{2}$.
The arguments of Chapter \ref{Section-BootAtInfty}, to establish statements {\bf I2}, proceed in three stages.
\noindent {\it Control of $\norm{u}_{H^3}$}
We explicitly calculate $\frac{d}{dt}\norm{\grad^3u}_{L^2}^2$ and seek to estimate the resulting error terms separately in two regions of space.
First, away from the singularity, on the truly three-dimensional region that includes the origin. Here the estimates are simpler, due to hypotheses {\bf H2.2} and {\bf H2.3}.
Second, on a toroidal region that includes the singular ring. This region requires more delicacy, and we split the solution into the rescaled almost self-similar profile, and $\widetilde{u}$.
Since $\Qb$ is smooth, the higher-order norms scale exactly with $\frac{1}{\lambda}$. In particular,
\begin{equation}\label{Strategy2-scalingQb}
\norm{\frac{1}{\lambda}\Qb(y)}_{H^3(\real^3)} \leq \frac{C(\Qb)}{\lambda^3(t)},
\end{equation}
where the constant is uniform for all $b$ sufficiently small - see Lemma \ref{Lemma-UnprovenProperty}. Note that equation (\ref{Strategy2-scalingQb}) is better than {\bf H2.1}.
For terms in $\widetilde{u}$, note that the $H^1$ norm is better than $\frac{1}{\lambda}$ due to {\bf H1.3}. By assuming $m>0$ is sufficiently small, we use this superior $H^1$ control to offset our use of {\bf H2.1}. We prove that,
\[
\abs{\frac{d}{dt}\norm{u}_{H^3}^2} \lesssim \frac{1}{\lambda^8} + \frac{e^{-\frac{\sigma_4}{b}}}{\lambda^2}\norm{u}_{H^3}^2.
\]
To prove {\bf I2.1}, we use equation (\ref{Strategy2-lambdaIntegralImproved}) to integrate carefully.
\noindent {\it Initial regularity improvement}
Let $\psi^A$ be a smooth cutoff function that covers the support of $\grad \chi$ - this is a toroidal shell that acts as an interface between the singular dynamics and the truly three-dimensional dynamics. We hope for any control of $\norm{\psi^A u}_{H^\nu}$ that is better than an interpolation of {\bf H2.1}. Calculate $\frac{d}{dt}\norm{\psi^Au}_{H^\nu}^2$ directly from equation (\ref{Eqn-NLS}) and integrate in time. The result is effectively Kato's smoothing effect and a Strichartz estimate,
\begin{equation}\label{Strategy2-KatoStrich}
\norm{\psi^A u}_{L^\infty_tH^\nu}^2
\lesssim \norm{\psi^B u}_{L^2_tH^{\nu+\frac{1}{2}}}^2
+ \int_0^t\abs{\int{D^\nu\left(\psi^Au\abs{u}^2\right)
\, D^\nu\left(\psi^A\overline{u}\right)}}.
\end{equation}
where $\psi^B$ is some other cutoff function with slightly larger support.
Due to equation (\ref{Strategy2-lambdaIntegralImproved}), we see that the term in $H^{\nu+\frac{1}{2}}$ is infact of the order $\frac{1}{\lambda^{2\left(\nu-\frac{1}{2}\right)}}$. This is exactly the sort of control we want, however the nonlinear term of (\ref{Strategy2-KatoStrich}) is uncooperative.
Since, $H^1(\real^2) \not\hookrightarrow L^\infty(\real^2)$, we cannot apply an $L^\infty$ norm without breaking scaling.
This is unlike the arguments of, \cite{R06, RaphaelSzeftel-StandingRingNDimQuintic-08}, where Strauss' radial interpolation inequality allows exactly such an embedding.
To estimate the nonlinear term of (\ref{Strategy2-KatoStrich}), we prove a modified Brezis-Gallou\"et estimate that does not break scaling `too-badly'. It is here that the form of hypothesis {\bf H2.1} is used delicately.
\noindent{\it Iterated Smoothing}
The next stage is to prove {\bf I2.2} and {\bf I2.3} hold on the support of $\grad \chi$. We iterate the argument of equation (\ref{Strategy2-KatoStrich}), in half-integer steps, beginning with $\nu = 3-\frac{1}{2}$, and introducing a new cutoff with smaller support each time. Due to the initial regularity improvement, it is possible to handle the nonlinear term of (\ref{Strategy2-KatoStrich}) systematically, and at the same order as the term in $H^{\nu+\frac{1}{2}}$. Due to integration (\ref{Strategy2-lambdaIntegralImproved}), at each stage we may smooth (almost) a half-derivative farther, relative to scaling, than was proved in the previous stage. After three iterates, we find that $\norm{\psi^C u}_{H^\frac{3}{2}}$ is (almost) order-zero in $\frac{1}{\lambda}$. The final iterate proves $\norm{\psi^D u}_{H^1}$ is constant.
To complete the proof of {\bf I2.2} and {\bf I2.3}, we repeat the iteration scheme for $\chi u$. The combination of hypotheses {\bf H2.2} and {\bf H2.3} with the results of the first iteration make the second iteration substantially simpler.
\begin{comment}
\begin{remark}\label{Remark-BoundAtOriginIsOverkill}
To complement Proposition \ref{Prop-Improv1}, we only need to bootstrap a bound for $\norm{\chi u}_{H^\frac{3}{2}}$, and in this sense (\ref{Hypo2-Hcrit}) is overkill.
Nevertheless, this hypothesis will allow a systematic proof for Lemma \ref{Lemma-NKSobolev} below. We do not investigate whether one might relax from $H^{3}$ data to hypothesize in a larger space, such as $H^{2}$.
\end{remark}
\end{comment}
\section{Proof of Log-log Singular Behaviour}
\label{Section-BootLoglog}
In this chapter we will prove that properties {\bf I1.1} through {\bf I1.5} are a consequence of hypotheses {\bf H1.1} through {\bf H1.5} and the bound,
\begin{equation}\label{Hypo2-HcritReinterpreted}
\norm{\chi u(t)}_{H^\frac{1}{2}} < \left(\alpha^*\right)^\frac{1}{10},
\end{equation}
which is a particular consequence of {\bf H2.3}.
\subsection{Almost Self-similar Profiles}
\label{Subsec-SelfSim}
Forthcoming parameter $\eta > 0$ is universal, sufficiently small, and will be determined in Section \ref{Subsec-Lyapounov}. For $b \neq 0$ let,
\begin{equation}\label{DefnEqn-Rb}\begin{aligned}
R_b = \frac{2}{b}\sqrt{1-\eta} && \text{ and } && R_b^- = R_b\sqrt{1-\eta},
\end{aligned}\end{equation}
and let $\phi_b$ denote a radially symmetric cutoff function with,
\begin{equation}\label{DefnEqn-phib}\begin{aligned}
\phi_b(y) = \left\{\begin{aligned}
1 && \text{ for } &\abs{y} \leq R_b^-\\
0 && \text{ for } &\abs{y} \geq R_b
\end{aligned}\right.
&& \text{ and }
&& \abs{\grad\phi_b}_{L^\infty} + \abs{\laplacian\phi_b}_{L^\infty} \to 0 \text{ as } \abs{b} \to 0.
\end{aligned}\end{equation}
The following result was original shown in \cite[Prop 1]{MR-SharpUpperL2Critical-03}. The refined cutoff, with parameter $\eta$, is introduced in \cite[Prop 8 and 9]{MR-UniversalityBlowupL2Critical-04}.
\begin{prop}[Localized Self-Similar Profiles]\label{Prop-SelfSimilar}
For all $\eta > 0$ sufficiently small there exists positive $b^*(\eta)$ and $\delta(\eta)$ such that for all $\abs{b} < b^*(\eta)$ there exists a unique radial solution $Q_b$ to,
\begin{equation}\label{DefnEqn-Qb}
\left\{\begin{aligned}
&\laplacian Q_b - Q_b + ib\Lambda Q_b + Q_b\abs{Q_b}^2 = 0,\\
&\begin{aligned}
P_b = Q_b e^{i\frac{b\abs{y}^2}{4}} > 0
&& \text{ for } y\in[0,R_b),
\end{aligned}\\
&\begin{aligned}
\abs{Q_b(0)-Q(0)} < \delta(\eta),
&& Q_b(R_b) = 0.
\end{aligned}
\end{aligned}\right.
\end{equation}
The truncation to $\abs{y} < \frac{2}{b}$, $\widetilde{Q}_b(y) = Q_b(y)\phi_b(y)$, satisfies,
\begin{equation}\label{DefnEqn-TildeQb}
\laplacian\widetilde{Q}_b -\widetilde{Q}_b + ib\Lambda\widetilde{Q}_b + \widetilde{Q}_b\abs{\widetilde{Q}_b}^2 = - \Psi_b,
\end{equation}
with the explicit error term,
\begin{equation}\label{DefnEqn-Psib}
-\Psi_b = Q_b\laplacian\phi_b + 2\grad\phi_b\cdot\grad Q_b + ibQ_by\cdot\grad\phi_b+\left(\phi_b^3 - \phi_b\right)Q_b\abs{Q_b}^2.
\end{equation}
Moreover, $\widetilde{Q}_b$ satisfies the following properties:
\begin{itemize}
\item {\it Uniform closeness to the ground state}:
\begin{equation}\label{Eqn-Qb_closeToQ}\begin{aligned}
\norm{e^{C\abs{y}}\left(\widetilde{Q}_b-Q\right)}_{C^{3}} \rightarrow 0
&& \text{ as }
&& b \rightarrow 0.
\end{aligned}\end{equation}
\item {\it Derivative with respect to $b$}:
\begin{equation}\label{Eqn-Qb_by_b}\begin{aligned}
\norm{e^{C\abs{y}}
\left(\frac{\partial}{\partial b}\widetilde{Q}_b + i\frac{\abs{y}^2}{4}Q\right)}_{C^2} \rightarrow 0
&& \text{ as }
&& b \rightarrow 0.
\end{aligned}\end{equation}
\item {\it Supercritical mass}:
\begin{equation}\label{Eqn-Qb_by_bSquared}\begin{aligned}
\left. \frac{d}{d(b^2)}\left(\int{\abs{\widetilde{Q}_b}^2}\right)\right|_{b^2=0} = d_0
&& \text{ with }
&& 0 < d_0 < +\infty.
\end{aligned}\end{equation}
\end{itemize}
As a consequence of (\ref{Eqn-Qb_closeToQ}), for any polynomial $P(y)$ and $k=0,1$,
\begin{equation}\label{Eqn-PsibEstimate}
\abs{P(y)\grad^k\Psi_b}_{L^\infty} \leq e^{-\frac{C(P)}{\abs{b}}}.
\end{equation}
In particular, energy and momentum are degenerate,
\begin{equation}\label{Eqn-Qb_enerAndMoment}\begin{aligned}
\abs{E\left(\widetilde{Q}_b\right)} \leq e^{-(1-C\eta)\frac{\pi}{\abs{b}}}
&& \text{ and }
&& Im\left(\int{\grad_y\widetilde{Q}_b\,\overline{\widetilde{Q}_b}}\right) = 0.
\end{aligned}\end{equation}
\end{prop}
The linearized Schr\"odinger operator near $\widetilde{Q}_b$ is,
$M\begin{bmatrix}v\\iw\end{bmatrix} = M_+(v,w) + iM_-(v,w)$,
with,
\begin{align}\label{DefnEqn-M}
M_+(v,w) =&
-\laplacian_y v + v
-\left(\frac{\widetilde{Q}_b^2}{\abs{\widetilde{Q}_b}^2}+2\right)
\abs{\widetilde{Q}_b}^2v
-Im(\widetilde{Q}_b^2)\,w, \\
M_-(v,w) =&
-\laplacian_y w + w
-\left(2-\frac{\widetilde{Q}_b^2}{\abs{\widetilde{Q}_b}^2}\right)
\abs{\widetilde{Q}_b}^2w
-Im(\widetilde{Q}_b^2)\,v.
\end{align}
As with $L$, equation (\ref{DefnEqn-L}), there is an associated bilinear operator,
\begin{equation}\label{DefnEqn-Hb}
H_b(\epsilon,\epsilon) = H(\epsilon,\epsilon) + \widetilde{H}_b(\epsilon,\epsilon),
\end{equation}
where $H(\epsilon,\epsilon)$ is the usual form (\ref{DefnEqn-H}) associated with $L$. The correction term may be written,
\begin{equation}\label{DefnEqn-HbTilde}
\widetilde{H}_b(\epsilon,\epsilon) = \int{V_{11}\epsRe^2}+\int{V_{12}\epsRe\epsIm}+\int{V_{22}\epsIm^2},
\end{equation}
for well-localized potentials built on $\widetilde{Q}_b$, $Q$ and $y\cdot\grad$ - see \cite[Appendix C]{MR-UniversalityBlowupL2Critical-04}. Due to proximity with $Q$, equation (\ref{Eqn-Qb_closeToQ}), there is universal constant $C$ with,
\begin{equation}\label{Eqn-HbTildeEst}\begin{aligned}
\norm{e^{C\abs{y}}V_{ij}}_{L^\infty} \rightarrow 0
&& \text{ as }
&& b \rightarrow 0.
\end{aligned}\end{equation}
The following variation of $H$ is of a different nature. Let,
\begin{equation}\label{DefnEqn-HTilde}
\widetilde{H}(\epsilon,\epsilon) =
H(\epsilon,\epsilon) - \frac{1}{\norm{\Lambda Q}_{L^2}^2}\left(\epsRe,L_+\Lambda^2Q\right)\left(\epsRe,\Lambda Q\right),
\end{equation}
which simply alters the definition of ${\mathcal L}_+$, (\ref{DefnEqn-H}).
The following is a consequence of equation $(\ref{Strategy1-Algebra})$ and the Spectral Property,
\begin{lemma}[Alternative Spectral Property, {\cite[page 616]{MR-UniversalityBlowupL2Critical-04}}]
\label{Lemma-AlternateSpectral}
There exists a universal constant $\delta_0 > \tilde{\delta}_0 > 0$ such that $\forall \epsilon \in H^1$:
\begin{equation}\label{Eqn-2DSpectral-Alternate}\begin{aligned}
\widetilde{H}(\epsilon,\epsilon) \geq &\tilde{\delta}_0
\left(\int_{y\in\real^2}{\abs{\grad_y\epsilon}^2} +
\int_{y\in\real^2}{\abs{\epsilon^2}e^{-\abs{y}}}\right)\\
&-\frac{1}{\tilde{\delta}_0}\left(
\begin{aligned}&Re\left(\epsilon,Q\right) + Re\left(\epsilon,\abs{y}^2 Q\right) +
Re\left(\epsilon,yQ\right)\\ &+ Im\left(\epsilon,\Lambda Q\right) +
Im\left(\epsilon,\Lambda^2Q\right) + Im\left(\epsilon,\grad Q\right)\end{aligned}
\right)^2.
\end{aligned}\end{equation}
\end{lemma}
\begin{comment}
Neglecting the nonlinear term of (\ref{DefnEqn-Qb}), we propagate the truncation error $\Psi_b$ to get the linear radiation $\zeta_b$, itself truncated to be $\widetilde{\zeta}_b$. In section \ref{Subsec-Lyapounov} we will find the profile $\widetilde{Q}_b+\widetilde{\zeta}_b$ gives an accurate description of mass ejection from the singular regime.
\end{comment}
The following result is proven \cite[Lemma 15]{MR-UniversalityBlowupL2Critical-04}. In Section \ref{Subsec-Lyapounov} we will find the study of linear radiation gives an accurate description of mass ejection from the singular regime.
\begin{lemma}[Linear Radiation]\label{Lemma-LinearRadiation}
There are universal constants $C>0$ and $\eta^*>0$ such that for all $0 < \eta < \eta^*$ there is $b^*(\eta) > 0$ such that for all $0 < b < b^*(\eta)$ the following is true.
There exists a unique radial solution $\zeta_b$ to:
\begin{equation}\label{DefnEqn-Zb}
\left\{\begin{aligned}
\laplacian\zeta_b - \zeta_b + ib\Lambda\zeta_b = \Psi_b,\\
\int{\abs{\grad\zeta_b}^2} < + \infty,
\end{aligned}\right.
\end{equation}
where $\Psi_b$ is the truncation error given by (\ref{DefnEqn-TildeQb}). Moreover, if we denote
\begin{equation}\label{DefnEqn-GammaB}
\Gamma_b = \lim_{\abs{y}\to+\infty}\abs{y}\abs{\zeta_b(y)}^2,
\end{equation}
then the solution satisfies:
\begin{itemize}
\item {\it Decay past the support of $\Psi_b$}:
\begin{equation}\begin{aligned}\label{Eqn-zb_H1L2Near}
\norm{\abs{y}\abs{\zeta_b}+\abs{y}^2\abs{\grad\zeta_b}}_{L^\infty(\abs{y}\geq R_b)}
&\leq \Gamma_b^{\frac{1}{2}-C\eta} < +\infty.
\end{aligned}\end{equation}
\item {\it Smallness in $\dot{H}^1$}:
\begin{equation}\label{Eqn-zb_smallH1}\begin{aligned}
\int{\abs{\grad_y\zeta_b}^2} \leq \Gamma_b^{1-C\eta}.
\end{aligned}\end{equation}
\item {\it Derivative with respect to $b$}:
\begin{equation}\label{Eqn-zb_by_b}
\norm{\frac{\partial\zeta_b}{\partial b}}_{C^1} \leq \Gamma_b^{\frac{1}{2}-C\eta}.
\end{equation}
\item {\it Stronger decay for larger $\abs{y}$}:
\begin{align}
\label{Eqn-zbH1Far}
&\norm{\abs{y}^2\abs{\grad\zeta_b}}_{L^\infty(\abs{y}\geq R_b^2)}
\leq C\frac{\Gamma_b^\frac{1}{2}}{\abs{b}}, \text{ and}\\
\label{Eqn-GammaBEstimate}
e^{-(1+C\eta)\frac{\pi}{b}}
\leq \frac{4}{5} \Gamma_b
\leq &\norm{\abs{y}^2\abs{\zeta_b}^2}_{L^\infty(\abs{y}\geq R_b^2)}
\leq e^{-(1-C\eta)\frac{\pi}{b}}.
\end{align}
As an estimate on $\Gamma_b$, (\ref{Eqn-GammaBEstimate}) will be indispensable.
\end{itemize}
\end{lemma}
Forthcoming parameter $a>0$ is universal, sufficiently small, will be determined in Section \ref{Subsec-Lyapounov}, and determines the choice of $\eta$.
We denote,
\begin{equation}\begin{aligned}
\label{DefnEqn-A}
A(t) = e^{a\frac{\pi}{b(t)}},
&& \text{ so that, }
&& \Gamma_b^{-\frac{a}{2}} \leq A \leq \Gamma_b^{-\frac{3a}{2}},
\end{aligned}\end{equation}
and let $\phi_A$ denote a radially symmetric cutoff function with,
\begin{equation}\begin{aligned}
\label{DefnEqn-phiA}
\phi_A(y) = \left\{\begin{aligned}
1 && \text{ for }\abs{y} \leq A\\
0 && \text{ for }\abs{y} \geq 2A.
\end{aligned}\right.
\end{aligned}\end{equation}
Truncated radiation $\widetilde{\zeta}_b(y)=\phi_A(y)\zeta_b$ satisfies:
\begin{equation}\label{DefnEqn-TildeZb}
\laplacian\widetilde{\zeta}_b - \widetilde{\zeta}_b +ib\Lambda\widetilde{\zeta}_b = \Psi_b + F,
\end{equation}
where the error term $F$ is explicit,
\begin{equation}\label{DefnEqn-F}
F = \zeta_b\laplacian\phi_A + 2\grad\phi_A\cdot\grad\zeta_b + ib\zeta_b y\cdot\grad\phi_A,
\end{equation}
and, in particular, by (\ref{Eqn-zbH1Far}) and (\ref{Eqn-GammaBEstimate}),
\begin{equation}\label{Eqn-FEstimate}
\abs{F}_{L^\infty}+\abs{y\cdot\grad F}_{L^\infty} \leq C\frac{\Gamma^\frac{1}{2}_b}{A}.
\end{equation}
\begin{remark
\label{Remark-ChoiceOf-a}
For smaller values of $\eta$ the central profiles $\widetilde{Q}_b$ approximate the mass of the singular region more closely, equation (\ref{DefnEqn-Rb}), at the cost that estimates (\ref{Eqn-Qb_closeToQ}) through (\ref{Eqn-Qb_enerAndMoment}) are only known for ever smaller values of $b$. When $\eta$ is larger, to compensate for the imperfection of our central profile we require more of the radiative tail to get an accurate picture of mass transport, requiring a larger choice of $a$.
See \cite[page 53]{MR-SharpLowerL2Critical-06} for similar remarks on the optimality in choice of $A(t)$.
\end{remark}
\subsection{Estimates directly due to Geometric Decomposition}
\label{Subsec-LocalVirial}
The following Lemma explains our choice of norm for $\epsilon$.
\begin{lemma}[Weighted and Local $L^2$ Estimates]
\label{Lemma-L2ByGrad}
For any $\kappa > 0$ and for all $v \in H^1(\real^2)$,
\begin{equation}\label{Eqn-ExpDecayByGrad}
\int_{y\in\real^2}{\abs{v(y)}^2e^{-\kappa\abs{y}}}
\leq C(\kappa)\left(\int{\abs{\grad v(y)}^2} + \int_{\abs{y} \leq 1}{\abs{v(y)}^2e^{-\abs{y}}}\right).
\end{equation}
\begin{equation}\label{Eqn-L2ByGrad}
\int_{\abs{y} \leq \kappa}{\abs{v(y)}^2} \leq C\,\kappa^2\log \kappa
\left(\int{\abs{\grad v(y)}^2}
+ \int_{\abs{y}\leq 1}{\abs{v(y)}^2e^{-\abs{y}}}\right).
\end{equation}
\end{lemma}
Equation (\ref{Eqn-L2ByGrad}) is due to {\cite[eqn (4.11)]{MR-SharpLowerL2Critical-06}}. While, the original proof of (\ref{Eqn-ExpDecayByGrad}), {\cite[Lemma 5]{MR-UniversalityBlowupL2Critical-04}}, has a flaw, the methods of \cite{MR-SharpLowerL2Critical-06} give an alternate proof.
\begin{remark}[Non-concern for $\mu$]
In practice, we apply these Lemmas and the interaction estimates below only on regions within $\{\abs{y} \lesssim A(t)\}$. That is, equation (\ref{Hypo1-consequenceForMu}) always applies and we may choose to include measure $\mu(y)$ as appropriate.
\end{remark}
\begin{comment}
\begin{proof}[Proof of (\ref{Eqn-ExpDecayByGrad}), Lemma \ref{Lemma-L2ByGrad}]
Suppose that $v\in C_0^\infty$. Then write $v(\rho,\theta)$ in terms of its Fourier series,
\[\begin{aligned}
v(\rho,\theta) = \sum_{k=-\infty}^\infty{v_k(\rho)\,e^{ik\theta}}
&& \text{ where }
&& v_k(\rho) = \frac{1}{2\pi}\int_0^{2\pi}{v(\rho,\theta)e^{-ik\theta}\,d\theta}.
\end{aligned}\]
For $k\neq 0$ we integrate by parts in $\theta$, use H\"older, and the form of $\grad$ in polar coordinates,
\[
\abs{v_k(\rho)} \leq \frac{1}{\abs{k}}\int_0^{2\pi}{\abs{\frac{\partial}{\partial\theta}v(\rho,\theta)}\,d\theta}
\leq C \frac{\rho}{\abs{k}}\left(\int_0^{2\pi}{\abs{\grad v(\rho,\theta)}^2\,d\theta}\right)^\frac{1}{2}.
\]
Then by integrating in $\real^2$ we obtain Hardy's Inequality for the nonradial part,
\begin{equation}\label{Proof-ExpDecayByGrad-Hardy}
\int_{\real^2}{\frac{\abs{v(y)-v_0(\abs{y})}^2}{\abs{y}^2}} \lesssim \int{\abs{\grad v(y)}^2},
\end{equation}
and clearly $e^{-\kappa\abs{y}} \leq C(\kappa)\frac{1}{\abs{y}^2}$. Concerning the radial component, choose some $\rho_0\in[\frac{1}{2},1]$ such that,
\[
\abs{v_0(\rho_0)}^2\leq \frac{2}{\pi}\int_{\frac{1}{2}}^1{\abs{v}\rho\,d\rho}
\leq C\int_{\abs{y}\leq 1}{\abs{v}^2e^{-\abs{y}}}.
\]
Then we may write $v_0(\rho) = v_0(\rho_0) + \int_{\rho_0}^\rho{\partial_\tau v_0(\tau)\,d\tau}$. With H\"older,
\begin{equation}\label{Proof-ExpDecayByGrad-RadialPart}
\abs{v_0(\rho)}^2 \lesssim \abs{v_0(\rho_0)}^2 +
\left(\int_{\rho_0}^\rho{\frac{1}{\tau}\,d\tau}\right)
\left(\int_{0}^\infty{\abs{\grad v_0(\tau)}^2\tau\,d\tau}\right).
\end{equation}
Multiply (\ref{Proof-ExpDecayByGrad-RadialPart}) by $e^{-\kappa\rho}\rho$ and integrate over $(\rho,\theta) \in [1,\infty)\times S^1$. With (\ref{Proof-ExpDecayByGrad-Hardy}) we have proven the desired bound for $\int_{\abs{y}\geq 1}{\abs{v}^2e^{-\kappa\abs{y}}}$. The remaining bound for $\int_{\abs{y} < 1}{\abs{v}^2e^{-\kappa\abs{y}}}$ is trivial.
\end{proof}
\end{comment}
\begin{lemma}[Estimates on Interaction Terms, {\cite[Section 5.3(C)]{MR-SharpUpperL2Critical-03}} ]\label{Lemma-InteractionEst}
For all $s\in[s_0,s_1)$,
\begin{itemize}
\item {\it First-Order Terms}
\begin{equation}\label{Eqn-1stOrderEst}
\abs{\left(\epsilon(y),P(y)\frac{d^k}{dy^k}\widetilde{Q}_b(y)\right)}
\leq C(P)\left(\epsNorm\right)^\frac{1}{2},
\end{equation}
where $P(y)$ is any polynomial and $0\leq k \leq 3$.
\item {\it Second-Order Terms}
\begin{equation}\label{Eqn-2ndOrderEst}
\abs{\left(R(\epsilon),P(y)\frac{d^k}{dy^k}\widetilde{Q}_b(y)\right)}
\leq C(P)\left(\epsNorm\right)
\end{equation}
where $P(y)$ is any polynomial, $0\leq k \leq 3$, and $R(\epsilon)$ is the terms of $(\epsilon+\widetilde{Q}_b)\abs{\epsilon+\widetilde{Q}_b}^2$ formally quadratic in $\epsilon$ - see equation (\ref{DefnEqn-R}).
\item {\it Localized Higher-Order Terms}
\begin{equation}\label{Eqn-3rdOrderEst-J}
\int{\abs{J(\epsilon)-\abs{\epsilon}^4}\mu(y)\,dy}
\leq \delta(\alpha^*)\left(\epsNorm\right),
\end{equation}
where $J(\epsilon)-\abs{\epsilon}^4 = 4\abs{\epsilon}^2Re\left(\epsilon\overline{\widetilde{Q}_b}\right)$, the term of $\abs{\epsilon+\widetilde{Q}_b}^4$ formally cubic in $\epsilon$ and localized to the support of $\widetilde{Q}_b$. Similarly,
\begin{equation}\label{Eqn-3rdOrderEst-RTilde}
\left(\widetilde{R}(\epsilon),\Lambda\widetilde{Q}_b\right)
\leq \delta(\alpha^*)\left(\epsNorm\right),
\end{equation}
where $\widetilde{R}(\epsilon) = \epsilon\abs{\epsilon}^2$, the term of $(\epsilon+\widetilde{Q}_b)\abs{\epsilon+\widetilde{Q}_b}^2$ formally cubic in $\epsilon$.
\end{itemize}
\end{lemma}
The following estimate is our first nontrivial departure from the $L^2$-critical argument.
\begin{lemma}[Complete Estimate on $J(\epsilon)$]
\label{Lemma-JEstimate}
For all $s\in[s_0,s_1)$,
\begin{equation}\label{Eqn-4thOrderEst}
\int{\abs{\epsilon(y)}^4\mu(y)\,dy} \leq \delta(\alpha^*)\left(\epsNorm\right).
\end{equation}
With equation (\ref{Eqn-3rdOrderEst-J}), this gives a complete estimate for $J(\epsilon)$.
\end{lemma}
\begin{proof}
Partition the support of $\epsilon$ into two and three dimensional regions,
\begin{equation}\label{Proof-4thOrder-eqn1}
\int{\abs{\epsilon(y)}^4\mu(y)\,dy}
= \int{\left(1-\chi^4\right)\abs{\epsilon(y)}^4\mu(y)\,dy}
+ \int{\abs{\chi\left(\lambda y+(r,z)(s)\right)\epsilon(y)}^4\mu(y)\,dy}.
\end{equation}
The first RH term is supported away from $r=0$, and due to {\bf H1.1} the support of $1-\chi^4$ is approximately, $\left\{\abs{y}<\frac{2}{3}\frac{1}{\lambda}\right\}$, so that, $\frac{1}{3} \lesssim \mu(y) \lesssim \frac{5}{3}$. We estimate this term by two-dimensional Sobolev embedding and the small mass assumption {\bf H1.2}.
Regarding the second RH term, the support of $\chi^4$ excludes the support of $\widetilde{Q}_b$ by the same reasons. Changing variables,
\begin{equation}\label{Proof-4thOrder-eqn2}
\int{\abs{\chi\left(x(y)\right)\epsilon(y)}^4\mu(y)\,dy}
=\lambda^2\int_{x\in\real^{3}}{\abs{\chi(x) u(x)}^4\,dx}.
\end{equation}
By the three-dimensional Sobolev embedding, $\dot{H}^{\frac{3}{4}} \hookrightarrow L^4(\real^{3})$, and interpolation,
\begin{equation}\begin{aligned}
\label{Proof-4thOrder-eqn3}
\lambda^2\int_{x\in\real^{3}}{\abs{\chi(x) u(x)}^4\,dx}
&\lesssim \norm{\chi u}_{\dot{H}^\frac{1}{2}}^2
\lambda^2\norm{\chi u}_{\dot{H}^1(\real^{3})}^2
&\lesssim \norm{\chi u}_{H^\frac{1}{2}}\left(\int{\abs{\grad_y\epsilon}^2\mu\,dy}\right).
\end{aligned}\end{equation}
To complete the proof, we use the assumed control {\bf H2.3} for the first and only time.
\end{proof}
\begin{lemma}[Estimates due to Conservation Laws]
\label{Lemma-PrelimEstimatesConservLaws}
For all $s\in[s_0,s_1)$ the following are true:
\begin{itemize}
\item {\it Due to conservation of mass:}
\begin{equation}\label{Eqn-prelimMassEst}
b^2 + \int{\abs{\tilde{u}}^2} \lesssim \left(\alpha^*\right)^\frac{1}{2}.
\end{equation}
\item {\it Due to conservation of energy:}
\begin{multline}\label{Eqn-prelimEnerEst}
\abs{
2Re\left(\epsilon,\widetilde{Q}_b\right)
-\int{\abs{\grad\epsilon}^2\mu(y)\,dy}
+3\int_{\abs{y}\leq\frac{10}{b}}{Q^2\epsRe^2}
+\int_{\abs{y}\leq\frac{10}{b}}{Q^2\epsIm^2}
}\\
\shoveright{
\leq
\Gamma_b^{1-C\eta} + \delta(\alpha^*)\left(\epsNorm\right),
}
\end{multline}
\item {\it Due to localized momentum (\ref{Hypo1-enerMoment}):}
\begin{equation}\label{Eqn-prelimMomentEst}
\abs{Im\left(\epsilon,\grad\widetilde{Q}\right)}
\leq
\Gamma_b^2 + \delta(\alpha^*)\left(\epsNorm\right)^\frac{1}{2}.
\end{equation}
In particular, by H\"older and (\ref{Eqn-Qb_closeToQ}), (\ref{Eqn-prelimMomentEst}) also holds for $\abs{\left(\epsIm,Re(\grad\widetilde{Q}_b)\right)}$.
\end{itemize}
\end{lemma}
\begin{proof}
\begin{itemize}
\item {\it Conservation of mass}: $\int_{\real^{3}}{\abs{u(t)}^2\,dx} = \int{\abs{u_0}^2}$
From the geometric decomposition, expand and change some variables,
\begin{equation}\label{Proof-prelimMass-eqn1}
\int{\abs{\widetilde{Q}_b(y)}^2\mu(y)\,dy}
+2Re\left(\int{\epsilon\overline{\widetilde{Q}}_b\mu(y)\,dy}\right)
+\int{\abs{\tilde{u}(t)}^2}
=\int{\abs{u_0}^2}.
\end{equation}
Expand measure $\mu$.
Due to the bound on $\lambda$, equation (\ref{Hypo1-consequenceForLambdaGamma}), hypotheses {\bf H1.1} and {\bf H1.2}, and the supercritical mass of $\widetilde{Q}_b$,
\begin{equation}\label{Proof-prelimMass-eqn2}\begin{aligned}
\int{\abs{\widetilde{Q}_b}^2\mu(y)\,dy} - \int{Q^2} =
& \lambda\int{y_1\abs{\widetilde{Q}_b}^2\,dy}
+(r(t)-1)\int{\abs{\widetilde{Q}_b}^2}\\
& +\int{\abs{\widetilde{Q}_b}^2} - \int{Q^2}
\gtrsim b^2 - \sqrt{\alpha^*}.
\end{aligned}\end{equation}
Due to small $b_0$ and the small mass of $\epsilon_0$, {\bf C1.2}
$\abs{\int_{\real^{3}}{\abs{u_0}^2}-\int_{\real^2}{Q^2}} \lesssim C\alpha^*$.
Due to local support, and hypothesis {\bf H1.2},
$
\abs{\int{\epsilon\overline{\widetilde{Q}}_b\mu}} \lesssim \alpha^*.
$
\item {\it Conservation of Energy}: $\int_{\real^{3}}{\abs{\grad u(t)}^2\,dx}-\frac{1}{2}\int{\abs{u}^4} = 2E_0$
From the geometric decomposition,
\begin{equation}\label{Proof-prelimEnergy-eqn1}
2\lambda^2E_0 = \int{\abs{\grad_y(\widetilde{Q}_b+\epsilon)}^2\mu(y)\,dy} - \frac{1}{2}\int{\abs{\widetilde{Q}_b+\epsilon}^4\mu(y)\,dy}
\end{equation}
Partially expand measure $\mu$,
\begin{equation}\label{Proof-prelimEnergy-eqn2}\begin{aligned}
\int{\abs{\grad_y(\widetilde{Q}_b+\epsilon)}^2\mu(y)\,dy}
= &
r(t)\int{\abs{\grad_y\widetilde{Q}_b}^2}
+ 2r(t)Re\left(\int{\grad_y\epsilon\cdot\grad_y\overline{\widetilde{Q}_b}}\right)
+ \int{\abs{\grad_y\epsilon}^2\mu(y)\,dy} \\
&
+\int{\lambda y_1
\left(\abs{\grad_y\widetilde{Q}_b}^2
+ 2Re(\epsilon\overline{\widetilde{Q}_b})\right)\,dy}.
\end{aligned}\end{equation}
Due to the support of $\widetilde{Q}_b$, the second line is of the order $\lambda$, and thus inconsequential.
Through a similar approach,
\begin{equation}\label{Proof-prelimEnergy-eqn3}\begin{aligned}
-\frac{1}{2}\int{\abs{\widetilde{Q}_b+\epsilon}^4\mu(y)\,dy}
= -&r(t)\left(\begin{aligned}
&\frac{1}{2}\int{\abs{\widetilde{Q}_b}^4}
+2Re\left(\int{\epsilon\overline{\widetilde{Q}_b}\abs{\widetilde{Q}_b}^2}\right)\\
&+\int{\abs{\epsilon}^2\abs{\widetilde{Q}_b}^2}
+Re\left(\int{\epsilon^2\overline{\widetilde{Q}_b}^2}\right)
\end{aligned}\right)\\
&+\lambda\, O\left(\abs{\widetilde{Q}_b}^2\right)
-\frac{1}{2}\int{J(\epsilon)\mu(y)\,dy}.
\end{aligned}\end{equation}
Now proceed as in the $L^2$-critical argument. Integrate $\int{\grad_y\epsilon\cdot\grad_y\overline{\widetilde{Q}_b}}$ by parts and substitute the equation for $\widetilde{Q}_b$ (\ref{DefnEqn-TildeQb}) - this cancels the term of (\ref{Proof-prelimEnergy-eqn3}) linear in $\epsilon$. Recall the bound for $\Psi_b$ (\ref{Eqn-PsibEstimate}), the degenerate energy of $\widetilde{Q}_b$ (\ref{Eqn-Qb_enerAndMoment}), proximity to $Q$ (\ref{Eqn-Qb_closeToQ}), that $r(t) \sim 1$, and the non-trivial estimate on $J$, equation (\ref{Eqn-4thOrderEst}).
\item {\it Localized momentum (\ref{Hypo1-enerMoment})}:
In cylindrical coordinates, $\grad_xf\cdot\grad_xg = \partial_rf\partial_rg + \partial_zf\partial_zf$. For this proof we denote $r$ by $x_1$ and $z$ by $x_2$. Fix either $j=1$ or $j=2$.
From the geometric decomposition,
\begin{multline}\label{Proof-prelimMoment-eqn1}
\lambda(t)Im\left(\int_{\real^{3}}{\partial_{x_j}\psi^{(x)}\partial_{x_j} u\,\overline{u}\,dx}\right)
=
Im\left(\int{
\partial_{x_j}\psi^{(x)
\partial_{y_j}\left(\widetilde{Q}_b+\epsilon\right)\,
\overline{\left(\widetilde{Q}_b+\epsilon\right)}\mu(y)\,dy}\right).
\end{multline}
Directly from definition (\ref{DefnEqn-psi-x}), $\partial_{x_j}\psi^{(x)} = 1$ on the support of $\widetilde{Q}_b$.
For the interaction term in $\partial_{y_j}\epsilon\overline{\widetilde{Q}_b}$ we expand the measure $\mu$ (\ref{DefnEqn-mu}) and integrate by parts the term in $r(t)$. Applying the denegerate momentum of $\widetilde{Q}_b$ (\ref{Eqn-Qb_enerAndMoment}) we have,
\begin{equation}\label{Proof-prelimMoment-eqn2}\begin{aligned}
2r(t)Im\left(\epsilon,\partial_{y_j}\widetilde{Q}_b\right) =
& Im\left(\int{\lambda y_1
\left(\partial_{y_j}\epsilon\overline{\widetilde{Q}_b} + \partial_{y_j}\widetilde{Q}_b\overline{\epsilon} + \partial_{y_j}\widetilde{Q}_b\overline{\widetilde{Q}_b}\right)\,dy}\right)\\
&+Im\left(\int{\partial_{x_j}\psi^{(x)
\partial_{y_j}\epsilon\overline{\epsilon}\mu(y)\,dy}\right)\\
&-\lambda(t)Im\left(\int_{\real^{3}}{\partial_{x_j}\psi^{(x)}\partial_{x_j} u\,\overline{u}\,dx}\right).
\end{aligned}\end{equation}
The first line is the order $\lambda$, and thus negligible.
The second line we apply H\"older and the small mass assumption {\bf H1.2}. The final term is controlled by {\bf H1.4}.
\begin{remark}[Role of Momentum Conservation]
The estimate analogous to (\ref{Eqn-prelimMomentEst}) in the $L^2$-critical context is proven with the conservation of momentum in place of {\bf H1.4}, \cite[Appendix A]{MR-SharpUpperL2Critical-03}.
As might be expected, the proof of {\bf I1.4} will resemble the proof of momentum conservation. See equation (\ref{Proof-Improv1-enerMoment-MorawetzType}).
\end{remark}
\end{itemize}
\end{proof}
\begin{definition}[NLS Reformulated for $\epsilon$]
\label{Defn-epsEqn}
For $s\in[s_0,s_1)$, $y \in \left\lbrack-\frac{r(t)}{\lambda(t)},+\infty\right)\times\real$, and a suitable boundary condition at $y_1 = -\frac{r(t)}{\lambda(t)}$, $\epsilon$ satisfies:
\begin{equation}\begin{aligned}
\label{DefnEqn-epsEqn}
ib_s\frac{\partial\widetilde{Q}_b}{\partial b}
+i\epsilon_s
-M(\epsilon) + \frac{\lambda}{r(y_1)}\partial_{y_1}\epsilon
+ib\Lambda\epsilon
= &
i\left(\frac{\lambda_s}{\lambda}+b\right)\Lambda\widetilde{Q}_b
+\tilde{\gamma}_s\widetilde{Q}_b
+i\frac{(r_s,z_s)}{\lambda}\cdot\grad_y\widetilde{Q}_b\\
&
+i\left(\frac{\lambda_s}{\lambda}+b\right)\Lambda\epsilon
+\tilde{\gamma}_s\epsilon
+i\frac{(r_s,z_s)}{\lambda}\cdot\grad_y\epsilon\\
&
+\Psi_b
-R(\epsilon),
\end{aligned}\end{equation}
where we introduced the new variable,
\begin{equation}\label{DefnEqn-gammaTilde}
\tilde{\gamma}(s) = -s - \gamma(s).
\end{equation}
Note the single new term due to cylindrical symmetry.
As already mentioned, the term $R(\epsilon)$ corresponds to those terms formally quadratic in $\epsilon$,
\begin{equation}\label{DefnEqn-R}
R(\epsilon) =
\left(\epsilon+\widetilde{Q}_b\right)\abs{\epsilon+\widetilde{Q}_b}^2
-\widetilde{Q}_b\abs{\widetilde{Q}_b}^2
-2\abs{\widetilde{Q}_b}^2\epsilon - \left(2\widetilde{Q}_b^2-Re(\widetilde{Q}_b^2)\right)\overline{\epsilon}.
\end{equation}
\begin{comment}
R(\epsilon) =
\left(\epsilon+\widetilde{Q}_b\right)\abs{\epsilon+\widetilde{Q}_b}^2
-\widetilde{Q}_b\abs{\widetilde{Q}_b}^2
-\left((1-i)\widetilde{Q}_b^2 + (1+i)\abs{\widetilde{Q}_b}^2 +1\right)\epsilon
-iIm(\widetilde{Q}_b^2)\overline{\epsilon}.
\end{comment}
\end{definition}
\begin{lemma}[Estimates due to Orthogonality Conditions]
\label{Lemma-PrelimEstimatesOrthogConds}
For all $s\in[s_0,s_1)$,
\begin{equation}\label{Eqn-prelimLambda+BEst}
\abs{\frac{\lambda_s}{\lambda}+b} +\abs{b_s}
\lesssim
\Gamma_b^{1-C\eta} + \left(\epsNorm\right),
\end{equation}
\begin{multline}\label{Eqn-prelimGamma+REst}
\abs{\tilde{\gamma}_s -
\frac{1}{\abs{\Lambda Q}_{L^2}^2}\left(\epsRe,L_+(\Lambda^2Q)\right)}
+ \abs{\frac{r_s}{\lambda}}+\abs{\frac{z_s}{\lambda}}\\
{
\leq
\Gamma_b^{1-C\eta} + \delta(\alpha^*)\left(\epsNorm\right)^\frac{1}{2}.
}
\end{multline}
\end{lemma}
Estimates (\ref{Eqn-prelimLambda+BEst}) and (\ref{Eqn-prelimGamma+REst}) are a direct result of orthogonality conditions (\ref{Eqn-GeoDecomp-OrthogReY2}), (\ref{Eqn-GeoDecomp-OrthogReY}), (\ref{Eqn-GeoDecomp-OrthogImL2}) and (\ref{Eqn-GeoDecomp-OrthogImL}) by taking the respective inner products with the $\epsilon$ equation (\ref{DefnEqn-epsEqn}). Due to equation (\ref{Hypo1-consequenceForLambdaGamma}), terms resulting from $\frac{\lambda}{r(\lambda)}\partial_{y_1}\epsilon$ are inconsequential.
The estimates due to energy and momentum, (\ref{Eqn-prelimEnerEst}) and (\ref{Eqn-prelimMomentEst}), are involved in the estimates of $\abs{b_s}$ and $\abs{\frac{r_s}{\lambda}}+\abs{\frac{z_s}{\lambda}}$ respectively.
Otherwise, all calculations are localized to the support of $\widetilde{Q}_b$ and are identical to the $L^2$-critical argument.
See \cite[Appendix C]{MR-UniversalityBlowupL2Critical-04} or \cite[Appendix A]{R-StabilityOfLogLog-05} for the complete calculations.
\begin{lemma}[Local Virial Identity]\label{Lemma-LocalVirial}
For all $s\in[s_0,s_1)$,
\begin{equation}\label{Eqn-LocalVirial}
b_s \geq \delta_1\left(\epsNorm\right) - \Gamma_b^{1-C\eta},
\end{equation}
where $\delta_1 > 0$ is a universal constant.
\end{lemma}
\begin{proof}[Brief Proof]
Begin with the method used to prove preliminary estimate (\ref{Eqn-prelimLambda+BEst}). Take the real part of the inner product of $\epsilon$ equation (\ref{DefnEqn-epsEqn}) with $\Lambda\widetilde{Q}_b$. Recognize that $\partial_sIm\left(\epsilon,\Lambda\widetilde{Q}_b\right) = 0$ due to orthogonality condition (\ref{Eqn-GeoDecomp-OrthogImL}).
An adapted version of the algebraic property $L_+(\Lambda Q) = -2Q$ is applied, \cite[equation (101)]{MR-UniversalityBlowupL2Critical-04}.
After recognizing the equation of $\widetilde{Q}_b$,
injecting the conservation of energy cancels the remaining terms linear in $\epsilon$.
The resulting terms quadratic in $\epsilon$ are the bilinear operator $H_b(\epsilon,\epsilon)$, equation (\ref{DefnEqn-Hb}). The remaining terms cubic in $\epsilon$ (due to the original inner product) were estimated as part of Lemma \ref{Lemma-InteractionEst}.
See \cite[Appendix C]{MR-UniversalityBlowupL2Critical-04}
for the complete calculation. Controlling the auxilliary terms of the conservation of energy with equation (\ref{Eqn-prelimEnerEst}) we have,
\begin{equation}\label{Proof-LocalVirial-firstMain}\begin{aligned}
-b_s\,Im\left(\frac{\partial}{\partial b}\widetilde{Q}_b,\Lambda\widetilde{Q}_b\right)
\gtrsim
&H_b(\epsilon,\epsilon)\\
&+b_s\,Im\left(\epsilon,\Lambda\frac{\partial}{\partial b}\widetilde{Q}_b\right)
-\left(\frac{\lambda_s}{\lambda} + b\right)
Im\left(\epsilon,\Lambda^2\widetilde{Q}_b\right)\\
&-\tilde{\gamma}_s\,Re\left(\epsilon,\Lambda\widetilde{Q}_b\right)
-\frac{(r_s,z_s)}{\lambda}\cdot Im\left(\epsilon,\grad\widetilde{Q}_b\right)\\
&-\Gamma_b^{1-C\eta} - \delta(\alpha^*)\left(\epsNorm\right).
\end{aligned}\end{equation}
Recall that $\partial_b\widetilde{Q}_b \approx -i\frac{\abs{y}^2}{4}Q$,
make the correction (\ref{Eqn-HbTildeEst}) for $\widetilde{H}_b$,
and apply preliminary estimates (\ref{Eqn-prelimLambda+BEst}) and (\ref{Eqn-prelimGamma+REst}). With the proximity to $Q$ we may write,
\begin{multline}\label{Proof-LocalVirial-secondMain}
b_s\,\frac{1}{4}\norm{yQ}_{L^2}^2
\gtrsim
H(\epsilon,\epsilon) - \tilde{\gamma}_s\left(\epsRe,\Lambda Q\right)\\
-\Gamma_b^{1-C\eta} - \delta(\alpha^*)\left(\epsNorm\right).
\end{multline}
Identify the alternate form of $\widetilde{H}$, equation (\ref{DefnEqn-HTilde}), apply the preliminary estimate for $\tilde{\gamma}_s$, equation (\ref{Eqn-prelimGamma+REst}), and apply the adapted version of the Spectral Property, Lemma \ref{Lemma-AlternateSpectral}.
\end{proof}
\begin{remark}[Progress in proving Proposition \ref{Prop-Improv}]
\label{Remark-waypointToImprov1}
We have already proven the first half of {\bf I1.2} as the preliminary estimate (\ref{Eqn-prelimMassEst}). The local virial identity with preliminary estimate (\ref{Eqn-prelimLambda+BEst}) produce a closed expression for $\lambda$ and $b$, which we treat with simple arguments to prove the following Lemma. In particular,
equation (\ref{Eqn-lambdaUpperBound}) implies the
1\textsuperscript{st} lower bound of {\bf I1.3}.
Following similar methods, we will then prove, the 2\textsuperscript{nd} upper bound of
{\bf I1.3}, {\bf I1.4}, {\bf I1.5} and {\bf I1.1}.
\end{remark}
\begin{lemma}[Upper Bound on Blowup Rate]\label{Lemma-Upperbound}
For all $s\in[s_0,s_1)$,
\begin{equation}\label{Eqn-bLowerBound}
b(s) \geq \frac{3\pi}{4\log s}, \text{ and}
\end{equation}
\begin{equation}\label{Eqn-lambdaUpperBound}
\lambda(s) \leq \sqrt{\lambda_0}e^{-\frac{\pi}{3}\frac{s}{\log s}}.
\end{equation}
\end{lemma}
\begin{proof}
Inject hypothesis {\bf H1.2} into the local virial identity (\ref{Eqn-LocalVirial}) and carefully integrate in time. From $b>0$, the bound on $\Gamma_b$ (\ref{Eqn-GammaBEstimate}), and the clever choice of $s_0$, equation (\ref{DefnEqn-s}),
\begin{equation}\label{Proof-UpperBound-eqn1}\begin{aligned}
\partial_se^{+\frac{3\pi}{4b}} = -\frac{b_s}{b^2}\frac{3\pi}{4}e^{+\frac{3\pi}{4b}} \leq 1
&& \Longrightarrow
&& e^{+\frac{3\pi}{4b}} \leq s -s_0 + e^{+\frac{3\pi}{4b_0}} = s.
\end{aligned}\end{equation}
This proves (\ref{Eqn-bLowerBound}).
Next, we view preliminary estimate (\ref{Eqn-prelimLambda+BEst}) and hypothesis {\bf H1.2} as the approximate dynamics of $\lambda$,
\begin{equation}\label{Eqn-lambda-prelimDynamics}
\abs{\frac{\lambda_s}{\lambda}+b}+\abs{b_s} < \Gamma_b^\frac{1}{2}.
\end{equation}
In particular as $b>0$ is small, $-\frac{\lambda_s}{\lambda} \geq \frac{2b}{3}$, which we integrate with (\ref{Eqn-bLowerBound}),
\begin{equation}\label{Proof-UpperBound-eqn2}
-\log\lambda \geq -\log\lambda_0 + \int_{s_0}^{s}{\frac{\pi}{2\log\sigma}\,d\sigma}.
\end{equation}
Assume $s_0$ is sufficiently large through choice of data (\ref{DataP1-mass}) with $\alpha^*$ sufficiently small, then,
\begin{equation}\label{Proof-UpperBound-eqn3}
\int_{s_0}^s{\frac{\pi}{2\log\sigma}d\sigma}
\geq \frac{\pi}{3}\left(\frac{s}{\log s}-\frac{s_0}{\log s_0}\right).
\end{equation}
From the choice of data {\bf C1.3},
and (\ref{DefnEqn-s}), $-\log\lambda_0 \geq e^{\frac{\pi}{2b_0}} = s_0^\frac{3}{2}$. Thus we have proven (\ref{Eqn-lambdaUpperBound}),
\
-\log\lambda \geq -\frac{1}{2}\log\lambda_0 + \frac{\pi}{3}\frac{s}{\log s}.
\]
\end{proof}
\begin{proof}[Corollary of (\ref{Eqn-lambdaUpperBound})]
By simple change of variables, (\ref{Eqn-lambdaUpperBound}), and choice of data (\ref{DataP1-mass}) and (\ref{DataP1-loglog}),
\begin{equation}\label{Eqn-t1Small}
T_{hyp}
= \int_{s_0}^{s_1}{\lambda^2(\sigma)\,d\sigma}
\leq \lambda_0\int_{2}^{+\infty}{e^{-\frac{2\pi}{3}\frac{s}{\log s}}\,ds}
< \alpha^*.
\end{equation}
\end{proof}
\begin{proof}[Proof of {\bf I1.3}, 2\textsuperscript{nd} upper bound]
As a direct consequence of (\ref{Eqn-lambdaUpperBound}), again assuming $s_0 > 0$ sufficiently large,
\begin{equation}\label{Proof-Improv1-loglog-eqn1}
-\log\left(s\lambda(s)\right) \geq \frac{\pi}{3}\frac{s}{\log s} - \log s \geq \frac{s}{\log s}.
\end{equation}
Taking the logarithm and applying equation (\ref{Eqn-bLowerBound}),
\begin{equation}\label{Proof-Improv1-loglog-eqn2}\begin{aligned}
\log\abs{-\log\left(s\lambda(s)\right)}
\geq \log\left(\frac{s}{\log s}\right)
\geq \frac{4}{15}\log s \geq \frac{\pi}{5 b(s)}
&& \Longrightarrow
&& s\lambda(s) \leq e^{-e^{\frac{\pi}{5b}}},
\end{aligned}\end{equation}
which in particular implies $\lambda \leq e^{-e^{\frac{\pi}{5b}}}$, the second upper bound of {\bf I1.3}
\end{proof}
\begin{proof}[Proof of {\bf I1.4}]
Recall approximate dynamic (\ref{Eqn-lambda-prelimDynamics}), which was due to preliminary estimate (\ref{Eqn-prelimLambda+BEst}) and the hypothesized control on $\epsilon$. As a consequence, for $s\in[s_0,s_1)$,
\begin{equation}\label{Proof-Improv1-enerMoment-eqn1}\begin{aligned}
\frac{d}{ds}\left(\lambda^2e^\frac{5\pi}{b}\right)
= 2\lambda^2e^\frac{5\pi}{b}\left(\frac{\lambda_s}{\lambda}+b-b-\frac{5\pi b_s}{2b^2}\right)
\leq& -\lambda^2be^{5\pi}{b} < 0,\\
&\Longrightarrow \lambda^2(t)e^{\frac{5\pi}{b(t)}} \leq \lambda_0^2e^\frac{5\pi}{b_0}.
\end{aligned}\end{equation}
Then, with the estimate on $\Gamma_b$ (\ref{Eqn-GammaBEstimate}), the choice of data (\ref{DataP1-enerMoment}) and the estimate on $\Gamma_b$ again,
\begin{equation}\label{Proof-Improv1-enerMoment-eqn2}
\lambda^2(t)\abs{E_0} < \Gamma_{b(t)}^4\,e^\frac{5\pi}{b_0}\lambda_0^2\abs{E_0}
<\Gamma_{b(t)}^4\,e^\frac{5\pi}{b_0}\Gamma_{b_0}^{10} \ll \Gamma_{b(t)}^4,
\end{equation}
which proves the energy-normalization part of {\bf I1.4}. Regarding the localized momentum, calculate directly from equation (\ref{Eqn-NLS}) that,
\begin{equation}\begin{aligned}\label{Proof-Improv1-enerMoment-MorawetzType}
\frac{d}{dt}Im\left(\int{\grad\psi^{(x)}\cdot\grad u\overline{u}}\right)
= &
Re\left(\int{\partial_{x_j}\partial_{x_k}\psi^{(x)}
\partial_{x_k}u\partial_{x_j}\overline{u}}\right)\\
&-\frac{1}{2}\int{\laplacian\psi^{(x)}\abs{u}^4}
-\frac{1}{4}\int{\laplacian^2\psi^{(x)}\abs{u}^2}.
\end{aligned}\end{equation}
This is a special case of the general Morawetz calculaton - eg. \cite[equation (3.36)]{Tao06}.
Recall from definition (\ref{DefnEqn-psi-x}) that the support of $\psi^{(x)}$ is well away from $r=0$. Apply the two-dimensional Sobolev embedding $H^\frac{1}{2} \hookrightarrow L^4$ to estimate,
\begin{equation}\label{Proof-Improv1-enerMoment-eqn4}
\abs{\frac{d}{dt}Im\left(\int{\grad\psi^{(x)}\cdot\grad u\overline{u}}\right)}
\leq C(\psi^{(x)})\norm{u(t)}_{H^1}^2 \lesssim \frac{1}{\lambda^2},
\end{equation}
where the final inequality is due to hypothesized control on $\epsilon$ and the small excess mass {\bf H1.2}. Note that, $\int_0^t{\frac{d\tau}{\lambda^2(\tau)}} = \int_{s_0}^s{d\sigma} \leq s$, and so we have proven,
\begin{equation}\label{Proof-Improv1-enerMoment-eqn5}
\lambda(t)\abs{Im\left(\grad\psi^{(x)}\cdot\grad u(t)\overline{u}(t)\right)}
\leq \lambda(t)\abs{Im\left(\grad\psi^{(x)}\cdot\grad u_0\overline{u}_0\right)}
+ C\lambda(t)s(t).
\end{equation}
Due to the estimate on $\Gamma_b$ (\ref{Eqn-GammaBEstimate}) and equation (\ref{Proof-Improv1-loglog-eqn2}) from the previous proof, $C\lambda(t)s(t) \leq C\Gamma_{b(t)}^{10} \ll \Gamma_b^4$.
Then by virtually the same calculation as equations (\ref{Proof-Improv1-enerMoment-eqn1}) and (\ref{Proof-Improv1-enerMoment-eqn2}), for $s\in[s_0,s_1)$,
\begin{equation}\label{Proof-Improv1-enerMoment-eqn6}
\frac{d}{ds}\left(\lambda e^\frac{6\pi}{b}\right) \leq -\frac{1}{2}\lambda be^\frac{6\pi}{b} < 0
\Longrightarrow \lambda(t) e^\frac{6\pi}{b(t)} \leq \lambda_0e^\frac{6\pi}{b_0},
\end{equation}
and so by the estimate on $\Gamma_b$ and choice of data (\ref{DataP1-enerMoment}),
\[
\lambda(t)\abs{Im\left(\grad\psi^{(x)}\cdot\grad u_0\overline{u}_0\right)}
\leq \Gamma_{b(t)}^5\,e^\frac{6\pi}{b_0}\Gamma_{b_0}^{10} \ll \Gamma_{b(t)}^4.
\]
This proves the localized-momentum part of {\bf I1.4}.
\end{proof}
\begin{proof}[Proof of {\bf I1.5}]
We follow the argument found in the proof of \cite[Lemma 7]{R-StabilityOfLogLog-05}.
Fix some $s_2 \leq s_3 \in [s_0,s_1)$.
Substitute the local virial identity (\ref{Eqn-LocalVirial}) into the preliminary estimate (\ref{Eqn-prelimLambda+BEst}) to control the norm of $\epsilon$. With a crude bound for $\Gamma_b$,
\begin{equation}\label{Proof-Improv1-almostMono-eqn1}
\abs{\frac{\lambda_s}{\lambda} + b} \leq C \left( b_s + b^2 \right),
\end{equation}
From hypothesis {\bf H1.2}, $0 < b^2 < \delta(\alpha^*)b$
where $\delta(\alpha^*) \rightarrow 0$ as $\alpha^* \to 0$. Then,
\begin{equation}\label{Proof-Improv1-almostMono-eqn2}
-\log\left(\frac{\lambda(s_2)}{\lambda(s_3)}\right)
= \int_{s_2}^{s_3}{\left(\frac{\lambda_s}{\lambda}+b\right)} - \int_{s_2}^{s_3}{b}
\leq \delta(\alpha^*) - \frac{1}{2}\int_{s_2}^{s_3}{b}
\leq \delta(\alpha^*).
\end{equation}
In particular, we may assume $\alpha^*$ is such that $\delta(\alpha^*) < \log 2$, which proves {\bf H1.5}.
\end{proof}
\begin{proof}[Proof of {\bf H1.1}]
Preliminary estimate (\ref{Eqn-prelimGamma+REst}) can be crudely simplified to,
\begin{equation}\label{Proof-Improv1-rz-eqn1}
\abs{\frac{r_s}{\lambda}}+\abs{\frac{z_s}{\lambda}} \leq 1.
\end{equation}
Then we have for all $s\in[s_0,s_1)$,
\begin{equation}\label{Proof-Improv1-rz-eqn2}
\abs{r(s)-r_0}+\abs{z(s)-z_0}
\leq \int_{s_0}^s{\abs{r_s}+\abs{z_s}}
\leq \int_{s_0}^s{\lambda(\sigma)\,d\sigma}
\leq \sqrt{\lambda_0}\int_2^{+\infty}{e^{-\frac{\pi}{3}\frac{\sigma}{\log\sigma}}\,d\sigma}
<\alpha^*,
\end{equation}
where we applied (\ref{Eqn-lambdaUpperBound}), the choice of data (\ref{DataP1-loglog}) and the smallness of $b_0$ (\ref{DataP1-mass}). With our choice of $r_0$,$z_0$ (\ref{DataP1-rz}), this proves {\bf H1.1}.
\end{proof}
\subsection{Lyapounov Functional}
\label{Subsec-Lyapounov}
\begin{comment}
The remaining challenge is to establish the improved bound for $\epsilon$ (\ref{Improv1-epsilon}). Where parameter $b$ is monotone decreasing, this control is a trivial consequence of the local virial identity (\ref{Eqn-LocalVirial}). However, this may not be for all times, so we will establish a Lyapounov functional as an alternate source of monotinicity.
The centerpiece observation will be the following: accounting for precisely the excess mass at the singularity
above the mass of $Q$, the Lyapounov functional is bounded above and below by $\epsTildeNorm$ to the order of $\Gamma_b^{1-Ca}$. This precise character allows us to bridge control for $\epsilon$ between times where $b$ is decreasing.
\end{comment}
To begin this section, we repeat the calculation of the local virial identity, this time including the linear radiation $\widetilde{\zeta}_b$ as part of the central profile. That is we write,
\begin{equation}\label{DefnEqn-epsTilde}\begin{aligned}
\tilde{\epsilon} = \epsilon - \widetilde{\zeta}_b
&& \Rightarrow
&& u(t,x) = \frac{1}{\lambda(t)}\left(\widetilde{Q}_{b(t)}+\widetilde{\zeta}_{b(t)}+\tilde{\epsilon}(t)\right)
\left(\frac{(r,z)-(r(t),z(t))}{\lambda}\right)e^{-i\gamma(t)},
\end{aligned}\end{equation}
where the parameters of the geometric decomposition are unchanged. The equation for $\tilde{\epsilon}$ may then be written analogous to (\ref{DefnEqn-epsEqn}), with a new linearized evolution operator analogous to $M$, (\ref{DefnEqn-M}).
\begin{lemma}[Radiative Virial Identity]\label{Lemma-RefinedVirial}
For all $s\in[s_0,s_1)$,
\begin{equation}\label{Eqn-RadiativeVirial}\begin{aligned}
\partial_s f_1 \geq &\delta_2\left(\epsTildeNorm\right)\\
&+\Gamma_b - \frac{1}{\delta_2}\int_{A\leq\abs{y}\leq 2A}{\abs{\epsilon}^2\,dy},
\end{aligned}\end{equation}
where $\delta_2, c > 0$ are universal constants and,
\begin{equation}\label{DefnEqn-f1}
f_1(s) =
\frac{b}{4}\abs{y\widetilde{Q}_b}_{L^2}^2
+ \frac{1}{2}Im\left(\int{y\cdot\grad\widetilde{\zeta}_b\overline{\widetilde{\zeta}}_b}\right)
+Im\left(\epsilon,\Lambda\widetilde{\zeta}_b\right).
\end{equation}
\end{lemma}
Compared with the local virial identity,
the radiative virial identity is useless to control $\epsilon$ in $\dot{H}^1$ due to the presence of mass term $\int_{A\leq\abs{y}\leq 2A}{\abs{\epsilon}^2}$.
See equation (\ref{Eqn-L2ByGrad}) for further discouragement. Nevertheless, we will link this term to the ejection of mass from the singularity, through the radiation, into the dispersive regime - Lemma \ref{Lemma-MassDisperse}. Then, we will show this mass ejection is more or less uninterrupted by demonstrating the Lyapounov fuctional - Lemma \ref{Lemma-LyapounovFunc}. Finally, through the conservation of energy we will prove precise bounds on the Lyapounov functional in terms of the excess mass at the singularity and $\abs{\epsilon}_{\dot{H}^1}$ - Lemma \ref{Lemma-LyapounovEstimates}. These bounds will allow us to bridge between times where $b_s \leq 0$ (times where the local virial identity is useful) to give a control for $\epsilon$ pointwise in time - Lemma \ref{Lemma-Lowerbound}.
Let $\phi_\infty$ be a smooth radial cutoff function on $\real^2$,
\begin{equation}\label{DefnEqn-phiInfty}\begin{aligned}
\phi_{\infty}(y) = \left\{\begin{aligned}
0 && \text{ for }\abs{y} \leq \frac{1}{2}\\
1 && \text{ for }\abs{y} \geq 3,
\end{aligned}\right.
&& \text{ and }
&& \begin{aligned}
\frac{1}{4}\leq\phi_{\infty}'\leq \frac{1}{2} &&& \text{ for }1\leq \abs{y}\leq 2,\\
0\leq \phi_{\infty}' &&& \text{ for all }y.
\end{aligned}
\end{aligned}\end{equation}
\begin{lemma}[Mass-ejection from Singular and Radiative Regimes]
\label{Lemma-MassDisperse}
\begin{equation}\label{Eqn-MassDisperse}
\partial_s\left(\frac{1}{r(t)}\int{\phi_{\infty}\left(\frac{y}{A}\right)\abs{\epsilon}^2\mu(y)\,dy}\right)
\geq \frac{b}{400}\int_{A\leq \abs{y} \leq 2A}{\abs{\epsilon}^2\,dy}
-\Gamma_b^\frac{a}{2}\int{\abs{\grad_y\epsilon}^2\mu(y)\,dy}
-\Gamma_b^2.
\end{equation}
\end{lemma}
\begin{remark}[Interpretation of Lemma \ref{Lemma-MassDisperse}]\label{Remark-MeaningOfMassDisperse}
Assume for the sake of heuristics that $\epsilon \approx \zeta_b$ on the region, $\abs{y} \sim A$. Then with the definition of $\Gamma_b$ (\ref{DefnEqn-GammaB}) and the assumed control on $\epsilon$ {\bf H1.2}, equation (\ref{Eqn-MassDisperse}) suggests continuous ejection of mass from the region $\abs{y} < \frac{A}{2}$, regardless of whether that region is growing or contracting.
\end{remark}
\begin{comment}
\begin{remark}[Role of $r_0 \sim 1$]
\label{Remark-RoleOfR_0=1}
In the coming argument we frequently refer to $r_0 \sim 1$, (\ref{DataP1-rz}). There is nothing intrinsic about radius one - the present argument works for any non-zero fixed radius.
Indeed, one should view the factor $\frac{1}{r(t)}$ (\ref{Eqn-MassDisperse}) as normalizing the measure $\mu$ in the singular region.
We will on occasion write proof in terms of $r_0$ to emphasize that $r_0 \sim 1$ is not fundamental. See, for example, (\ref{Proof-RefinedLyapounov-eqn5}).
\end{remark}
\end{comment}
\begin{lemma}[Lyapounov Functional, \cite{MR-SharpLowerL2Critical-06} ]\label{Lemma-LyapounovFunc}
For all $s\in[s_0,s_1)$,
\begin{equation}\label{Eqn-Lyapounov}
\partial_s{\mathcal J} \leq -Cb\left(
\Gamma_b + \epsTildeNorm + \int_{A\leq\abs{y}\leq 2A}{\abs{\epsilon}^2}
\right),
\end{equation}
where $C>0$ is a universal constant,
\begin{equation}\label{DefnEqn-LyapounovFunctional}\begin{aligned}
{\mathcal J}(s) =
&\int{\abs{\widetilde{Q}_b}^2} - \int{\abs{Q}^2}
+ 2Re\left(\epsilon,\widetilde{Q}_b\right)\\
&+\frac{1}{r(s)}\int{\left(1-\phi_{\infty}\left(\frac{y}{A}\right)\right)\abs{\epsilon}^2\mu(y)\,dy}\\
&-\frac{\delta_2}{800}\left(
b\tilde{f}_1(b) - \int_0^b{\tilde{f}_1(v)\,dv}
+b\,Im\left(\epsilon,\Lambda\widetilde{\zeta}_b\right) \right),
\end{aligned}\end{equation}
where $\tilde{f}_1$ is the principal part of $f_1$, equation (\ref{DefnEqn-f1}),
\begin{equation}\label{DefnEqn-f1Tilde}
\tilde{f}_1(b) =
\frac{b}{4}\abs{y\widetilde{Q}_b}_{L^2}^2
+ \frac{1}{2}Im\left(\int{y\cdot\grad\widetilde{\zeta}_b\overline{\widetilde{\zeta}}_b}\right).
\end{equation}
\end{lemma}
Lemma \ref{Lemma-LyapounovFunc} is proven from the radiative virial estimate (\ref{Eqn-RadiativeVirial}), mass dispersion estimate (\ref{Eqn-MassDisperse}), and the conservation of mass. The proof is provided at the end of the section.
Now let us discuss what ${\mathcal J}$ is.
\begin{lemma}[Estimates on Lyapounov Functional]\label{Lemma-LyapounovEstimates}
For all $s\in[s_0,s_1)$ we have the crude estimate,
\begin{equation}\label{Eqn-CrudeLyapounovEst}
\abs{ {\mathcal J} - d_0b^2 } < \delta_3b^2,
\end{equation}
where $0 < \delta_3 \ll 1$ is a universal constant, and $d_0b^2$ is the approximate excess mass of profile $\widetilde{Q}_b$ - see (\ref{Eqn-Qb_by_bSquared}). There also holds a more refined estimate,
\begin{equation}\label{Eqn-RefinedLyapounovEst}
{\mathcal J}(s) - f_2(b(s)) \left\{\begin{aligned}
\leq
\Gamma_b^{1-Ca} + CA^2&\left(\epsNorm\right)\\
\geq
-\Gamma_b^{1-Ca} + \frac{1}{C}&\left(\epsNorm\right),
\end{aligned}\right.
\end{equation}
where $f_2$ is the principal part of ${\mathcal J}$ concerned with mass of the profile,
\begin{equation}\label{DefnEqn-f2}
f_2(b) =
\int{\abs{\widetilde{Q}_b}^2} - \int{\abs{Q}^2}
-\frac{\delta_2}{800}\left(
b\tilde{f}_1(b) - \int_0^b{\tilde{f}_1(v)\,dv} \right).
\end{equation}
\end{lemma}
\begin{proof
To prove (\ref{Eqn-CrudeLyapounovEst}) we will approximate each term of (\ref{DefnEqn-LyapounovFunctional}). To estimate the term in $\abs{\epsilon}^2$, recall the support of $\phi_\infty$ (\ref{DefnEqn-phiInfty}) and the consequence for $\mu(y)$, such as equation (\ref{Hypo1-consequenceForMu}). Then,
\begin{equation}\label{Proof-CrudeLyapounov-eqn1}\begin{aligned}
\int{\left(1-\phi_\infty\left(\frac{y}{A}\right)\right)\abs{\epsilon}^2\mu(y)\,dy}
&\lesssim \int_{\abs{y}\leq 3A}{\abs{\epsilon}^2}\\
&\lesssim A^2\log A\left(\epsNorm\right) \leq \Gamma_b^\frac{1}{2},
\end{aligned}\end{equation}
where the second inequality is due to Lemma \ref{Lemma-L2ByGrad}
and the final inequality is from the definition of $A$ (\ref{DefnEqn-A}) and the hypothesized control of $\epsilon$. Estimate $(\epsilon,\widetilde{Q}_b)$ by the same control, and the terms in $\widetilde{\zeta}_b$ by (\ref{Eqn-zb_smallH1}). Equation (\ref{Eqn-CrudeLyapounovEst}) then follows from (\ref{Eqn-Qb_by_bSquared}) by noting that the constant $\delta_2$ due to the radiative virial identity (\ref{Eqn-RadiativeVirial}) can be assumed small with respect to universal constant $d_0$, so that $0 < \left.\frac{\partial f_2}{\partial b^2}\right|_{b^2=0} < \infty$.
Next we prove the refined estimate. Note that,
\begin{equation}\label{Proof-RefinedLyapounov-eqn1}
{\mathcal J}(s) - f_2(b(s)) =
2Re\left(\epsilon,\widetilde{Q}_b\right)
+ \frac{1}{r(t)}\int{\left(1-\phi_\infty\right)\abs{\epsilon}^2\mu(y)}
- \frac{\delta_2}{800}b\,Im\left(\epsilon,\Lambda\widetilde{\zeta}_b\right).
\end{equation}
By the bounds for $\widetilde{\zeta}_b$,
Lemma \ref{Lemma-L2ByGrad},
and the choice of $A$,
\begin{equation}\label{Proof-RefinedLyapounov-eqn2}\begin{aligned}
\abs{Im\left(\epsilon,\Lambda\widetilde{\zeta}_b\right)}
&\leq \Gamma_b^{\frac{1}{2}-C\eta}
\left(\int_{\abs{y} \leq A}\abs{\epsilon}^2\right)^\frac{1}{2}\\
&\lesssim \Gamma_b^{\frac{1}{2}-C\eta}A\left(\log A\right)^\frac{1}{2}
\left( \epsNorm \right)^\frac{1}{2}\\
&\lesssim \Gamma_b^{1-Ca} + \left(\epsNorm\right).
\end{aligned}\end{equation}
Since $b$ is small, the contribution of (\ref{Proof-RefinedLyapounov-eqn2}) is a factor of $\alpha^*$ smaller than the desired bound. Similar terms will be omitted for the remainder of the proof.
Regarding the two other terms of (\ref{Proof-RefinedLyapounov-eqn1}), the term linear in $\epsilon$ we recognize from the conservation of energy (\ref{Eqn-prelimEnerEst}). Indeed, the upper bound for (\ref{Proof-RefinedLyapounov-eqn1}) follows from (\ref{Eqn-prelimEnerEst}) with (\ref{Eqn-ExpDecayByGrad}) and,
\begin{equation}\label{Proof-RefinedLyapounov-eqn3}
\int{\left(1-\phi_\infty\right)\abs{\epsilon}^2\mu(y)\,dy} \lesssim A^2\log A\left(\epsNorm\right)
\end{equation}
which is due to (\ref{Eqn-L2ByGrad}).
To establish a lower bound for (\ref{Proof-RefinedLyapounov-eqn1}) we will need the following Lemma - the proof is based on a spectral result due to \cite{MartelMerle01}, with additional properties proven \cite{Maris02} and \cite{McLeod93}. See \cite[Lemma 8]{MR-SharpLowerL2Critical-06} for that spectral property, and \cite[Appendix D]{MR-SharpLowerL2Critical-06} for proof of the Lemma.
\begin{lemma}[Elliptic estimate for $L$.]
\label{Lemma-EllipticEstimate}
Recall the linearized Schr\"odinger operator $L$ from (\ref{DefnEqn-L}).
There exists a universal constant $\delta_4 > 0$ such that $\forall v \in H^1(\real^2)$,
\begin{equation}\label{EllipticEstimate}\begin{aligned}
Re\left(L(v),v\right) - \int{\phi_\infty\abs{v}^2}
\geq &\delta_4\left(\int{\abs{\grad v}^2} + \int{\abs{v}^2e^{-\abs{y}}}\right)\\
&-\frac{1}{\delta_4}\left(
Re\left(v,Q\right) + Re\left(v,\abs{y}^2Q\right)
+ Re\left(v,yQ\right) + Im\left(v,\Lambda^2Q\right)
\right)^2.
\end{aligned}\end{equation}
\end{lemma}
Introduce a new radially symmetric cutoff function, analogous to $\phi_A$ (\ref{DefnEqn-phiA}) but with larger support such that $(1-\phi_B(y))(1-\phi_\infty(\frac{y}{A})) = 0$.
\begin{equation}\begin{aligned}
\label{DefnEqn-phiB}
\phi_B(y) = \left\{\begin{aligned}
1 && \text{ for }\abs{y} \leq 3A\\
0 && \text{ for }\abs{y} \geq 4A,
\end{aligned}\right.
\end{aligned}\end{equation}
From equation (\ref{Hypo1-consequenceForMu}), we may rewrite the principal part of the conservation of energy estimate (\ref{Eqn-prelimEnerEst}) as,
\begin{equation}\label{Proof-RefinedLyapounov-eqn4}\begin{aligned}
2Re\left(\epsilon,\widetilde{Q}_b\right) \approx&
\int{\left(1-\phi_B^2\right)\abs{\grad\epsilon}^2\mu(y)\,dy}\\
&+\int{\phi_B^2\abs{\grad\epsilon}^2\,dy}
-3\int{Q^2(\phi_B\epsRe)^2} - \int{Q^2(\phi_B\epsIm)^2},
\end{aligned}\end{equation}
where we used the exponential spatial decay of $Q$ and the lower bound for $\Gamma_b$ (\ref{DefnEqn-GammaB}) to control the excess in $Q^2\epsilon^2$ on $\abs{y} > \frac{10}{b}$.
With integration by parts,
\begin{equation}\label{Proof-RefinedLyapounov-eqn5.5}\begin{aligned}
\int{\phi_B^2\abs{\grad\epsilon}^2\,dy}
= \int{\abs{\grad(\phi_B\epsilon)}^2\,dy}
+\int{\laplacian\phi_B\,\phi_B\abs{\epsilon}^2\,dy}.
\end{aligned}\end{equation}
The principal part of (\ref{Proof-RefinedLyapounov-eqn1}) is then,
\begin{equation}\label{Proof-RefinedLyapounov-eqn7}\begin{aligned}
2Re\left(\epsilon,\widetilde{Q}_b\right)
&+ \frac{1}{r(t)}\int{(1-\phi_\infty)\abs{\epsilon}^2\mu(y)\,dy}\\
\approx&
\int{\left(1-\phi_B^2\right)\abs{\grad\epsilon}^2\mu(y)\,dy}\\
&+\left( Re\left(L(\phi_B\epsilon),\phi_B\epsilon\right)
- \int{\phi_\infty\abs{\phi_B\epsilon}^2} \right)
+ \int{\laplacian\phi_B\,\phi_B\abs{\epsilon}^2}\\
&+\int{ \left(1-\phi_\infty\right)\left(\frac{\mu}{r(t)}-\phi_B^2\right)
\abs{\epsilon}^2}.
\end{aligned}\end{equation}
The final term can be neglected, as $\left(1-\phi_\infty\right)\left(\frac{\mu}{r(t)}-\phi_B^2\right)$ is of the order $\lambda y_1$, and supported on $\abs{y} < 4A$.
The lower bound for (\ref{Proof-RefinedLyapounov-eqn1}) then follows from Lemma \ref{Lemma-EllipticEstimate}, an integration by parts, and the straightforward comparison,
\begin{equation}\label{Proof-RefinedLyapounov-eqn8}
\int{\phi_B^2\abs{\grad\epsilon}^2} + \int{\abs{\phi_B\epsilon}^2e^{-\abs{y}}}
\gtrsim \int{\phi_B^2\abs{\grad\epsilon}^2\mu(y)} + \int_{\abs{y}\leq\frac{10}{b}}{\abs{\epsilon}^2e^{-\abs{y}}},
\end{equation}
again due to the support of $\phi_B$ and the bound on $\lambda$.
This completes the proof of (\ref{Eqn-RefinedLyapounovEst}).
\end{proof}
\begin{lemma}[Lower Bound on Blowup Rate]\label{Lemma-Lowerbound}
For all $s\in[s_0,s_1)$,
\begin{equation}\label{Eqn-bUpperBound}
b(s) \leq \frac{4\pi}{3\log s},
\end{equation}
\begin{equation}\label{Eqn-epsIntegral}
\int_{s_0}^s{\left(
\Gamma_{b(\sigma)} + \epsNorm
\right)\,d\sigma}
\leq C \alpha^*,
\end{equation}
where $C > 0$ is a universal constant and,
\begin{equation}\label{Eqn-epsImproved}
\epsNorm \leq \Gamma_b^\frac{4}{5},
\end{equation}
which is equation (\ref{Improv1-epsilon}), the remaining part of {\bf I1.2}.
\end{lemma}
Note that (\ref{Eqn-bUpperBound}) is the 1\textsuperscript{st} upper bound of {\bf I1.3}. The only estimate remaining to establish Proposition \ref{Prop-Improv} follows as a corollary,
\begin{proof}[Proof of {\bf I1.3}, 2\textsuperscript{nd} lower bound]
Recall the approximate dynamics of $\lambda$, equation
(\ref{Eqn-lambda-prelimDynamics}). Since $b>0$ is small,
$-\frac{\lambda_s}{\lambda} \leq 3b$, which we integrate with (\ref{Eqn-bUpperBound}),
\begin{equation}\label{Proof-Improv1-loglog-eqn3}
-\log \lambda(s)
\leq -\log\lambda_0 + 4\pi\int_{s_0}^s{\frac{1}{\log\sigma}\,d\sigma}
\leq -\log\lambda_0 + 4\pi (s-s_0).
\end{equation}
Use (\ref{Eqn-bUpperBound}) again, and recall the definition of $s_0$ (\ref{DefnEqn-s}) and choice of data (\ref{DataP1-loglog}),
\begin{equation}\label{Eqn-lambdaLowerBound}
\lambda(s)
\geq \lambda_0e^{4\pi s_0}\,e^{-4\pi e^\frac{4\pi}{3b(s)}}
> e^{-e^\frac{5\pi}{b(s)}}.
\end{equation}
\end{proof}
\begin{proof}[Proof of Lemma \ref{Lemma-Lowerbound}]
To begin, note from the crude estimate (\ref{Eqn-CrudeLyapounovEst}) that we may divide the Lyapounov inequality (\ref{Eqn-Lyapounov}) by $\sqrt{\mathcal J}$ and integrate in time leaving,
\begin{equation}\label{Proof-LowerBound-eqn1}
\int_{s_0}^s{\left(
\Gamma_{b(\sigma)} + \epsNorm
\right)\,d\sigma}
\leq C\left(\sqrt{{\mathcal J}(s_0)} - \sqrt{{\mathcal J}(s)}\right)
\leq Cb_0.
\end{equation}
The choice of data (\ref{DataP1-mass}) then proves (\ref{Eqn-epsIntegral}). Alternately, we may view the crude estimate (\ref{Eqn-CrudeLyapounovEst}) and the Lyapounov inequality (\ref{Eqn-Lyapounov}) as giving a differential inequality for ${\mathcal J}$,
\begin{equation}\label{Proof-LowerBound-eqn2}\begin{aligned}
\partial_se^{+\frac{5\pi}{4}\sqrt{\frac{d_0}{{\mathcal J}}}}
\gtrsim \frac{b}{{\mathcal J}}\,\Gamma_be^{\frac{5\pi}{4}\sqrt{\frac{d_0}{{\mathcal J}}}}
\geq 1
&& \Longrightarrow
&& e^{+\frac{5\pi}{4}\sqrt{\frac{d_0}{{\mathcal J}(s)}}} \geq
e^{+\frac{5\pi}{4}\sqrt{\frac{d_0}{{\mathcal J}(s_0)}}}
+s -s_0.
\end{aligned}\end{equation}
Note that here we applied the bound on $\Gamma_b$ (\ref{Eqn-GammaBEstimate}), for which it was essential $\frac{5}{4} > 1+C\eta$ - see Remark \ref{Remark-NormalBrezisFails}.
By crude estimate (\ref{Eqn-CrudeLyapounovEst}) and the definition of $s_0$ (\ref{DefnEqn-s}),
\begin{equation}\label{Proof-LowerBound-eqn2.5}
e^{+\frac{5\pi}{4}\sqrt{\frac{d_0}{{\mathcal J}(s_0)}}}
> e^\frac{\pi}{b_0} > s_0,
\end{equation}
which, again with estimate (\ref{Eqn-CrudeLyapounovEst}), proves (\ref{Eqn-bUpperBound}) from (\ref{Proof-LowerBound-eqn2}).
It remains to establish the pointwise control of $\epsilon$. Fix $s \in[s_0,s_1)$.
\begin{enumerate}
\item If $\partial_sb(s) \leq 0$, then (\ref{Eqn-epsImproved}) follows from the local virial identity, Lemma \ref{Lemma-LocalVirial}.
\item If $\partial_sb(s) > 0$, then there exists a largest interval $(s_+,s)$, with $s_0\leq s_+$, on which $\partial_sb >0$.
\
\begin{aligned}
\text{This implies, }
&& b(s_+) < b(s)
&& \text{ and either, }
&& \left.\begin{aligned}
\left(a\right) &&&s_+ = s_0, \\
\;\;\text{ or,}\\
\left(b\right) &&&\partial_sb(s_+) = 0.
\end{aligned}\right.
\end{aligned}\]
In case (a) or (b), by the choice of small $\epsilon_0$
or the local virial identity, respectively,
\
\int{\abs{\grad_y\epsilon(s_+,y)}^2\mu(y)\,dy}
+\int_{\abs{y}\leq\frac{10}{b(s_+)}}{\abs{\epsilon(s_+,y)}^2e^{-\abs{y}}\,dy}
\leq \Gamma_{b(s_+)}^\frac{6}{7}.
\]
From the upper bound of refined estimate (\ref{Eqn-RefinedLyapounovEst}), and assuming $a>0$ is sufficiently small,
\begin{equation}\label{Proof-LowerBound-eqn5}
{\mathcal J}(s_+) - f_2(b(s_+)) \leq \Gamma_{b(s_+)}^\frac{5}{6} < \Gamma_{b(s)}^\frac{5}{6}.
\end{equation}
Since ${\mathcal J}$ is non-increasing,
and from the lower bound of refined estimate (\ref{Eqn-RefinedLyapounovEst}),
\begin{equation}\label{Proof-LowerBound-eqn6}\begin{aligned}
\Gamma_{b(s)}^\frac{5}{6}
\geq &
{\mathcal J}(s) - f_2(b(s_+))\\
\gtrsim &
\left(\int{\abs{\grad_y\epsilon(s,y)}^2\mu(y)\,dy}
+\int_{\abs{y}\leq\frac{10}{b(s)}}{\abs{\epsilon(s,y)}^2e^{-\abs{y}}\,dy}
\right)\\
&\qquad-\Gamma_{b(s)}^{1-Ca} + \left(f_2(b(s)) - f_2(b(s_+))\right).
\end{aligned}\end{equation}
As noted in the proof of crude estimate (\ref{Eqn-CrudeLyapounovEst}), we may assume the constant $\delta_2$ of equation (\ref{Eqn-RadiativeVirial}) is sufficiently small relative to $d_0$, such that $0 < \left.\frac{\partial f_2}{\partial_{b^2}}\right|_{b^2=0} < \infty$, and proving that $\left(f_2(b(s)) - f_2(b(s_+))\right) > 0$. Assuming $a>0$ is sufficiently small, this proves (\ref{Eqn-epsImproved}).
\end{enumerate}
\end{proof}
\begin{proof}[Proof of Lemma \ref{Lemma-MassDisperse}, \cite{MR-SharpLowerL2Critical-06}]
Directly from equation (\ref{Eqn-NLS}),
\begin{multline}\label{Proof-MassDisperse-eqn1}
\frac{1}{2}\partial_s\left(\int{
\phi_{\infty}\left(\frac{(r,z) - (r(t),z(t))}{\lambda A}\right)
\abs{u}^2\,dx}\right)\\
\begin{aligned}=&
\frac{1}{\lambda A}Im\left(\int{\grad_x\phi_\infty\left(\frac{y}{A}\right)
\cdot\grad_x u\,\overline{u}\,dx}\right)\\
&-\frac{1}{2\lambda^2 A}\int{\left(
\left(\frac{\lambda_s}{\lambda} + \frac{A_s}{A}\right)
+\frac{\partial_s(r,z)}{\lambda}\right)
\cdot\grad_x\phi_\infty\left(\frac{y}{A}\right)
\abs{u}^2\,dx}.
\end{aligned}
\end{multline}
From the choice of $A$ (\ref{DefnEqn-A}) and $\phi_\infty$ (\ref{DefnEqn-phiInfty}), the support of $\widetilde{Q}_b$ and $\phi_\infty\left(\frac{y}{A}\right)$ are disjoint. With the geometric decomposition
and change of variables we may rewrite (\ref{Proof-MassDisperse-eqn1}) in terms of $\abs{\epsilon}^2$,
\begin{multline}\label{Proof-MassDisperse-eqn2}
\frac{1}{2}\frac{d}{ds}\int{\phi_\infty\left(\frac{y}{A}\right)\abs{\epsilon}^2\mu(y)\,dy}\\
\begin{aligned}
=&
\frac{1}{A}Im\left(\int{\grad_x\phi_\infty\left(\frac{y}{A}\right)\cdot\grad_y\epsilon\,\overline{\epsilon}\mu(y)\,dy}\right)
+
\frac{b}{2}\int{\frac{y}{A}\cdot\grad_x\phi_\infty\left(\frac{y}{A}\right)\abs{\epsilon}^2\mu(y)\,dy}\\
&-\frac{1}{2A}\int{\left(\left(\frac{\lambda_s}{\lambda}+b+\frac{A_s}{A}\right)y+\frac{\partial_s(r,z)}{\lambda}\right)
\cdot\grad_x\phi_\infty\left(\frac{y}{A}\right)\abs{\epsilon}^2\mu(y)\,dy}.
\end{aligned}
\end{multline}
By Cauchy-Schwarz, the definition of $A$ (\ref{DefnEqn-A}) and the lower bound on $\Gamma_b$ (\ref{Eqn-GammaBEstimate}),
\begin{multline}\label{Proof-MassDisperse-eqn3}
\abs{\frac{1}{A}Im\left(\int{\grad_x\phi_\infty\left(\frac{y}{A}\right)\cdot\grad_y\epsilon\,\overline{\epsilon}\mu(y)\,dy}\right)}\\
\begin{aligned}
\leq&\frac{1}{A}\left(\int{\abs{\grad\epsilon}^2\mu(y)\,dy}\right)^\frac{1}{2}
\left(\int{\abs{\grad_x\phi_\infty\left(\frac{y}{A}\right)}\abs{\epsilon}^2\mu(y)\,dy}\right)^\frac{1}{2}\\
\leq&\frac{1}{2}\Gamma_b^\frac{a}{2}\int{\abs{\grad\epsilon}^2\mu(y)\,dy}
+\frac{b}{40}\int{\abs{\grad_x\phi_\infty\left(\frac{y}{A}\right)}\abs{\epsilon}^2\mu(y)\,dy}.
\end{aligned}
\end{multline}
The factor $\frac{b}{40}$ is arbitrary by assuming $b$ is sufficiently small.
The following term is the principal part of (\ref{Proof-MassDisperse-eqn2}): from the support of ${\phi_\infty}'$, and that ${\phi_\infty}'\geq 0$, equation (\ref{DefnEqn-phiInfty}),
\begin{equation}\label{Proof-MassDisperse-eqn4}
\frac{b}{2}\int{\frac{y}{A}\cdot\grad_x\phi_\infty\left(\frac{y}{A}\right)\abs{\epsilon}^2\mu(y)\,dy}
\geq \frac{b}{5}\int{\abs{\grad_x\phi_\infty\left(\frac{y}{A}\right)}\abs{\epsilon}^2\mu(y)\,dy}.
\end{equation}
Regarding the last line of (\ref{Proof-MassDisperse-eqn2}), apply preliminary estimates (\ref{Eqn-prelimLambda+BEst}) and (\ref{Eqn-prelimGamma+REst}), the support of ${\phi_\infty}'$,
and the definition of $A$
to estimate,
\begin{equation}\label{Proof-MassDisperse-eqn5}
\abs{\frac{1}{2A}\left(\left(\frac{\lambda_s}{\lambda}+b+\frac{A_s}{A}\right)y+\frac{\partial_s(r,z)}{\lambda}\right)}
\leq \frac{b}{40}.
\end{equation}
Due to the bounds for ${\phi_\infty}'\left(\frac{y}{A}\right)$ on $A\leq\abs{y}\leq 2A$, and lower bounds for $\mu$ similar to equation (\ref{Hypo1-consequenceForMu}),
\begin{equation}\label{Proof-MassDisperse-eqn6}
\int{\abs{\grad_x\phi_\infty\left(\frac{y}{A}\right)}\abs{\epsilon}^2\mu(y)\,dy}
\geq \frac{1}{6}\int_{A \leq \abs{y} \leq 2A}{\abs{\epsilon}^2\,dy}.
\end{equation}
From (\ref{Proof-MassDisperse-eqn2}) we have proven,
\begin{equation}\label{Proof-MassDisperse-eqn7}
\frac{d}{ds}\int{\phi_\infty\left(\frac{y}{A}\right)\abs{\epsilon}^2\mu(y)\,dy}
\geq
\frac{b}{20}\int_{A \leq \abs{y} \leq 2A}{\abs{\epsilon}^2\,dy} -
\Gamma_b^\frac{a}{2}\int{\abs{\grad\epsilon}^2\mu(y)\,dy}.
\end{equation}
Finally note that by preliminary estimate (\ref{Eqn-prelimGamma+REst}), $r(t) \sim 1$ from {\bf H1.1}, change of variables, and the log-log relationship (\ref{Hypo1-consequenceForLambdaGamma}), we have the easy estimate,
\begin{equation}
\abs{\frac{r_s}{r^{2}(t)}\int{\phi_\infty\left(\frac{y}{A}\right)\abs{\epsilon}^2\mu(y)\,dy}}
\ll \lambda \int{\abs{\tilde{u}}^2} \ll \Gamma_b^2.
\end{equation}
This completes the proof of Lemma \ref{Lemma-MassDisperse}.
\end{proof}
\begin{proof}[Proof of Lemma \ref{Lemma-LyapounovFunc}]
Multiply the radiative virial identity (\ref{Eqn-RadiativeVirial}) by $\frac{\delta_2 b}{800}$, and sum with the mass ejection equation (\ref{Eqn-MassDisperse})
to cancel the bad sign of $\int_{A\leq\abs{y}\leq 2A}{\abs{\epsilon}^2}$,
\begin{multline}\label{Proof-LyapounovFunc-eqn1}
\partial_s\left(\frac{1}{r^k(t)}\int{\phi_{\infty}\left(\frac{y}{A}\right)\abs{\epsilon}^2\mu(y)\,dy}\right)
+ \frac{\delta_2b}{800}\partial_s f_1 \\
\begin{aligned}
\geq& \frac{\delta_2^2b}{800}\left(\epsTildeNorm\right)\\
&+\frac{b}{800}\int_{A\leq \abs{y} \leq 2A}{\abs{\epsilon}^2\,dy}
+\frac{\delta_2b}{1000}\Gamma_b
-\Gamma_b^\frac{a}{2}\int{\abs{\grad_y\epsilon}^2\mu(y)\,dy}.
\end{aligned}
\end{multline}
The final term of (\ref{Proof-LyapounovFunc-eqn1}) has the bad sign.
Recall $\epsilon = \widetilde{\epsilon} + \widetilde{\zeta}_b$, equation (\ref{DefnEqn-epsTilde}), and $\widetilde{\zeta}_b$ is small in $\dot{H}^1$, equation (\ref{Eqn-zb_smallH1}), with support on which we may estimate $\mu$ so that,
\begin{equation}\label{Proof-LyapounovFunc-eqn2}\begin{aligned}
\Gamma_b^\frac{a}{2}\int{\abs{\grad_y\epsilon}^2\mu(y)\,dy}
&\lesssim \Gamma_b^\frac{a}{2}\left(
\Gamma_b^{1-C\eta} + \int{\abs{\grad\widetilde{\epsilon}}^2\mu(y)\,dy}
\right)\\
&\leq \Gamma_b^{1+\frac{a}{4}}
+ \Gamma_b^\frac{a}{2}\int{\abs{\grad\widetilde{\epsilon}}^2\mu(y)\,dy},
\end{aligned}\end{equation}
where for the second inequality we require $a > 4C\eta$
- see Remark \ref{Remark-ChoiceOf-a}.
To rewrite $\frac{\delta_2b}{800}\partial_s f_1$, note that,
\begin{equation}\label{Proof-LyapounovFunc-eqn3}\begin{aligned}
b\partial_s f_1 = \partial_s\left(
b\tilde{f}_1(b) - \int_0^b{\tilde{f}_1(v)\,dv}
+b Im\left(\epsilon,\Lambda\widetilde{\zeta}_b\right)
\right)
- \partial_s b \,Im\left(\epsilon,\Lambda\widetilde{\zeta}_b\right),
\end{aligned}\end{equation}
where $\tilde{f}_1$ is the principal part of $f_1$, given by equations (\ref{DefnEqn-f1Tilde}) and (\ref{DefnEqn-f1}), respectively.
Estimate the final term of (\ref{Proof-LyapounovFunc-eqn3}) with a combination of preliminary estimate (\ref{Eqn-prelimLambda+BEst}), H\"older, Lemma \ref{Lemma-L2ByGrad}, and the hypothesis $\epsilon$ {\bf H1.2}.
Equation (\ref{Proof-LyapounovFunc-eqn1}) has transformed into,
\begin{multline}\label{Proof-LyapounovFunc-eqn4}
\partial_s\left(
\frac{1}{r(t)}\int{\phi_{\infty}\left(\frac{y}{A}\right)\abs{\epsilon}^2\mu(y)\,dy}
+
\frac{\delta_2}{800}\left(
b\tilde{f}_1(b) - \int_0^b{\tilde{f}_1(v)\,dv}
+b Im\left(\epsilon,\Lambda\widetilde{\zeta}_b\right)
\right)
\right)\\
\geq
\frac{\delta_2^2b}{800}\left(
\epsTildeNorm + \int_{A\leq \abs{y} \leq 2A}{\abs{\epsilon}^2\,dy}
\right)
+\frac{\delta_2b}{2000}\Gamma_b.
\end{multline}
To identity the LHS of (\ref{Proof-LyapounovFunc-eqn4}) with $-\partial_s{\mathcal J}$, inject the conservation of mass, $\int_{\real^{3}}{\abs{u(t)}^2} = \int{\abs{u_0}^2}$.
As we did before, equation (\ref{Proof-prelimMass-eqn1}), rewrite $u(t)$ with the geometric decomposition, expand the product, change variables, expand the measure $\mu$,
divide by $r(t)$, and take the derivative $\partial_s$,
\begin{comment}
\begin{multline}\label{Proof-LyapounovFunc-eqn5}
\int{\abs{\epsilon}^2\mu(y)\,dy}
+ r(t)\left(\int{\abs{\widetilde{Q}_b}^2} - \int{\abs{Q}^2}
+ 2Re\left(\epsilon,\widetilde{Q}_b\right)\right)\\
= \int{\abs{u_0}^2} - r(t)\int{\abs{Q}^2}
- \int{\lambda y_1\left(\abs{\widetilde{Q}_b}^2 +
2Re\left(\epsilon\overline{\widetilde{Q}_b}\right)\right)}.
\end{multline}
Then by direct calculation,
\end{comment}
\begin{equation}\label{Proof-LyapounovFunc-eqn6}\begin{aligned}
\partial_s\left(\frac{1}{r(t)}\int{\phi_{\infty}\left(\frac{y}{A}\right)\abs{\epsilon}^2\mu(y)\,dy}\right)
=&
-\partial_s\left(\int{\abs{\widetilde{Q}_b}^2} - \int{\abs{Q}^2}
+ 2Re\left(\epsilon,\widetilde{Q}_b\right)\right)\\
& -\partial_s\left(\frac{1}{r(t)}
\int{\lambda y_1P(\lambda y_1,r(t))\left(\abs{\widetilde{Q}_b}^2
+ 2Re\left(\epsilon\overline{\widetilde{Q}_b}\right)\right)}
\right)\\
&-\frac{\partial_sr}{r^{2}(t)}\int{\abs{u_0}^2}.
\end{aligned}\end{equation}
Through a combination of preliminary estimates (\ref{Eqn-prelimLambda+BEst}) and (\ref{Eqn-prelimGamma+REst}), the $\epsilon$-equation (\ref{DefnEqn-epsEqn}), and the log-log rate (\ref{Hypo1-consequenceForLambdaGamma}),
\[
\abs{-\partial_s\left(\frac{1}{r(t)}
\int{\lambda y_1P(\lambda y_1,r(t))\left(\abs{\widetilde{Q}_b}^2
+ 2Re\left(\epsilon\overline{\widetilde{Q}_b}\right)\right)}
\right)}
\lesssim \lambda < \Gamma_b^2.
\]
Likewise, $\abs{\frac{\partial_sr}{r^{2}(t)}}\int{\abs{u_0}^2} \lesssim \lambda\int{\abs{u_0}^2} < \Gamma_b^2$.
Inserting equation (\ref{Proof-LyapounovFunc-eqn6}) into (\ref{Proof-LyapounovFunc-eqn4}) completes the proof of Lemma \ref{Lemma-LyapounovFunc}.
\end{proof}
\section{Proof of Global Behaviour}
\label{Section-BootAtInfty}
In this chapter we prove that properties {\bf I2.1} through {\bf I2.3} are a consequence of hypotheses {\bf H1.1} through {\bf H2.3}. The following properties of the singular dynamic proven in Chapter \ref{Section-BootLoglog} will be used:
the specific log-log rate, the geometric decomposition and resulting control on $b_s$,
and the integrability of $\norm{\tilde{u}}_{L^2_tH^1_x}$.
\subsection{Growth of \texorpdfstring{$\norm{u}_{H^{3}}$ }{u in H\textthreesuperior}}
It is left until Chapter \ref{Section-FinalProof} to show that $\lambda^{-1}$ follows the log-log rate (\ref{Thm-MainResult-LogLog}). Here, we use the log-log rate in the form {\bf H1.3}, and the control of $b_s$, to prove directly that $\lambda^{-1}(t)$ has the same integrability in time as $\sqrt{\frac{\log\abs{\log(T-t)}}{T-t}}$.
The following Lemma was previously noted, \cite[equation (51)]{RaphaelSzeftel-StandingRingNDimQuintic-08}.
\begin{lemma}[Integrability due to log-log rate]\label{Lemma-lambdaIntegralCrude}
Let $0 \leq \mu <2$ and $\sigma_1 \in \real$. Then,
\begin{equation}\label{Eqn-crudeLambdaIntegral}
\int_0^t{\frac{e^{\frac{\sigma_1}{b(\tau)}}}{\lambda^\mu(\tau)}\,d\tau} \lesssim C(\mu,\sigma_1,\alpha^*),
\end{equation}
where for fixed $\mu$ and $\sigma_1$, $C(\mu,\sigma_1,\alpha^*)$ decays much faster than $e^{-\frac{1}{\alpha^*}}$ as $\alpha^* \rightarrow 0$.
\end{lemma}
\begin{proof}
From the log-log rate, {\bf H1.3},
\[\begin{aligned}
e^\frac{\pi}{10 b} < \abs{\log\lambda}
&&\text{ and, } &&
\frac{\pi}{10b} > \frac{1}{100}\log s && \Rightarrow && \frac{1}{\lambda} > e^{+e^{\frac{\pi}{10b}}} > e^{\left(s^\frac{1}{100}\right)}.
\end{aligned}\]
Then by change of variables
and almost-monotony of $\lambda$, {\bf H1.5},
\[
\int_0^t{\frac{e^{\frac{\sigma_1}{b(\tau)}}}{\lambda^\mu(\tau)}\,d\tau} < \int_{s_0}^s{\frac{\abs{\log\lambda}^\frac{10\sigma_1}{\pi}}{\lambda^{\mu-2}(\tau')}\,d\tau'}
\lesssim \frac{\abs{\log\lambda(t)}^\frac{10\sigma_1}{\pi}}{\lambda^{\mu-2}(t)}\left(s(t)-s_0\right)
\lesssim e^{(\mu-2)s^\frac{1}{100}(t)}.
\]
Finally, to prove the behaviour of $C(\mu,\sigma_1,\alpha^*)$, recall $s(t) \geq s_0 = e^\frac{3\pi}{4b_0}$, and $b_0 < \alpha^*$.
\end{proof}
\begin{remark}[Lemma \ref{Lemma-lambdaIntegralCrude} for $\mu \geq 2$]
From the log-log rate, {\bf H1.3},
$s(t)-s_0 \lesssim e^{\frac{10}{\pi}\frac{1}{b(t)}}$, so by the same proof,
\begin{equation}\label{Eqn-outdatedLambdaIntegral}
\int_0^t{\frac{1}{\lambda^\mu}} \lesssim \frac{e^{\frac{10}{\pi}\frac{1}{b(t)}}}{\lambda^{\mu-2}}.
\end{equation}
This is the primary integrability tool of \cite{RaphaelSzeftel-StandingRingNDimQuintic-08}. The following improvement will be crucial,
\end{remark}
\begin{lemma}[Refined Integrability due to control of $b_s$]\label{Lemma-lambdaIntegral}
Let $\mu > 2$, $\sigma^*$ arbitrary and assume $\alpha^*>0$ is sufficiently small. Then for any $\sigma_2 > 0$ and all $t\in[0,T_{hyp})$,
\begin{equation}\label{Eqn-lambdaIntegral}
\int_0^t{\frac{e^{-\frac{\sigma^*}{b(\tau)}}}{\lambda^\mu(\tau)}\,d\tau}
\leq
C(\mu,\sigma_2,\alpha^*)\,\frac{
e^{-\frac{\sigma^*}{b(t)}}e^{+\frac{\sigma_2}{b(t)}}
}{\lambda^{\mu-2}(t)},
\end{equation}
where, for fixed $\mu$ and $\sigma_2$, $C(\mu,\sigma_2,\alpha^*) \rightarrow 0$ as $\alpha^* \rightarrow 0$.
\end{lemma}
\begin{comment}
\begin{remark}
... that Lemma \ref{Lemma-lambdaIntegral} is strictly stronger than result due only to blowup rate. Show. Note (\ref{Hypo2-bs}) is a key ingredient in proving the log-log rate. Justify that $b_s$ estimate is natural, in light of Lemma \ref{Lemma-ControlledHnk}.
... how is this version of Lemma \ref{Lemma-lambdaIntegral} necessary?
... We will use Lemma \ref{Lemma-lambdaIntegral} to emphasize the overwhelming smallness due to $\norm{\tilde{u}}_{H^1}$.
\end{remark}
\end{comment}
\begin{proof}
To begin, we prove the case $\sigma^* = 0$. By direct calculation,
\[
\frac{d}{ds}\left(\frac{1}{b}\frac{1}{\lambda^{\mu-2}}\right)
=
\frac{1}{\lambda^{\mu-2}}\left((\mu-2) - \frac{b_s}{b^2} - (\mu-2)\frac{\frac{\lambda_s}{\lambda}+b}{b}\right).
\]
For $\alpha^*$ sufficiently small relative to $\mu$,
from {\bf H1.2}
and the control of $b_s$, (\ref{Eqn-prelimLambda+BEst}),
\begin{equation}\label{Proof-lambdaIntegral-eqn1}
\frac{1}{\lambda^\mu}
\leq
C(\mu)\frac{1}{\lambda^2}\,\frac{d}{ds}\left(\frac{1}{b}\frac{1}{\lambda^{\mu-2}}\right)
=
C(\mu)\frac{d}{dt}\left(\frac{1}{b}\frac{1}{\lambda^{\mu-2}}\right).
\end{equation}
After integration, we estimate $C(\mu)\frac{1}{b}\frac{1}{\lambda^{\mu-2}} \leq C(\mu,\sigma_2,\alpha^*)\frac{e^{+\frac{\sigma_2}{b}}}{\lambda^{\mu-2}}$.
For those cases where $\sigma^* \neq 0$, integrate by parts,
\begin{equation}\label{Proof-lambdaIntegral-eqn2}
\int_0^t{\frac{ e^{-\frac{\sigma^*}{b(\tau)}} }{\lambda^\mu(\tau)}d\,\tau}
=\left.
e^{-\frac{\sigma^*}{b(\tau)}}\int_0^\tau{\frac{1}{\lambda^\mu(\tau')}d\,\tau'}
\right|_0^t
-\int_0^t{\sigma^*\left(\frac{b_\tau}{b^2}e^{-\frac{\sigma^*}{b(\tau)}}\int_0^\tau{\frac{1}{\lambda^\mu(\tau')}d\,\tau'}\right)d\,\tau}.
\end{equation}
Apply the previous case to the first RH term. For the second term, make the change of variable $b_\tau = \frac{b_{s(\tau)}}{\lambda^2(\tau)}$ and apply the previous case for some $\sigma_2 << \frac{1}{2}$. Use (\ref{Eqn-prelimLambda+BEst}) to approximate $b_s$, and we have bound the second term by a small multiple of the LHS.
\end{proof}
Lemma \ref{Lemma-lambdaIntegral} is simply not true for $\mu = 2$. As a substitute, we prove a corollary of the integrated Lyapounov inequality, equation (\ref{Eqn-epsIntegral}).
\begin{corollary}\label{Corollary-lambdaIntegral2version2}
Let $\sigma_3 \geq 0$. Then for all $t\in[0,T_{hyp})$,
\begin{equation}\label{Eqn-lambdaIntegral2version2}
\int_0^t{e^{\frac{\sigma_3}{b(\tau)}}
\left(\norm{\tilde{u}(\tau)}_{H^1}^2 + \frac{\Gamma_{b(\tau)}}{\lambda^2(\tau)}\right)
\,d\tau}
\lesssim C(\alpha^*)e^{\frac{\sigma_3}{b(t)}}.
\end{equation}
\end{corollary}
\begin{proof}
By change of variables and integration by parts,
\[\begin{aligned}
\int_0^t{e^{\frac{\sigma_3}{b(\tau)}}
\left(\norm{\tilde{u}(\tau)}_{H^1}^2 + \frac{\Gamma_{b(\tau)}}{\lambda^2(\tau)}\right)
\,d\tau}
=&\left.e^{\frac{\sigma_3}{b(\sigma)}}
\int_0^\sigma{\lambda^2(\sigma ')\norm{\widetilde{u}(\sigma ')}_{H^1_x}^2
+ \Gamma_{b(\sigma ')}\,d\sigma '}
\right|_{s_0}^{s(t)}\\
&+\sigma_3\int_{s_0}^{s(t)}{\frac{b_s}{b^2}e^{\frac{\sigma_3}{b(\sigma)}}
\left(\int_0^\sigma{\lambda^2\norm{\widetilde{u}}_{H^1_x}^2+\Gamma_b}\right)\,d\sigma}.
\end{aligned}\]
Then observe the control on $b_s$ (\ref{Eqn-prelimLambda+BEst}) and estimate (\ref{Eqn-epsIntegral}).
\end{proof}
\begin{remark}[Optimality of (\ref{Eqn-lambdaIntegral2version2})
Corollary \ref{Corollary-lambdaIntegral2version2} is the best possible integrability of $\frac{e^\frac{\delta}{b}}{\lambda^2}$ for constant $\delta$.
As a heuristic, assume that $\lambda \sim \sqrt{T-t}$ and $e^{\frac{1}{b}} \sim \abs{\log \lambda} \sim \abs{\log(T-t)}$, motivated by the log-log rate {\bf H1.3}.
The integral, $\int^T{\frac{\abs{\log(T-t)}^{\delta}}{T-t}\,dt}$, is only finite for values of $\delta$ sufficiently negative. In our case, the maximum threshold for $\delta$ is given dynamically by (\ref{Eqn-epsIntegral}).
\end{remark}
Next, we translate hypotheses {\bf H2.1} through {\bf H2.3} into a gain of derivative during particular three-dimensional Sobolev embeddings. Consider a smooth cutoff function with support on $\chi^{-1}\{1\}$,
\begin{equation}\label{DefnEqn-TildeChi}
\widetilde{\chi}(r,z,\theta) = \left\{\begin{aligned}
1 && \text{ for } \abs{(r,z) - (1,0)} \geq \frac{3}{4}\\
0 && \text{ for } \abs{(r,z) - (1,0)} \leq \frac{2}{3}.
\end{aligned}\right.
\end{equation}
\begin{lemma}[Consequences of Bootstrap Hypotheses]\label{Lemma-NKSobolev}
Let $v = \widetilde{\chi} u$, and suppose that
$2 \geq l_1 \geq l_2 \geq l_3 \geq 0$ with $l_1+l_2+l_3 = 3$.
Then,
\begin{equation}\label{Eqn-NKSobolev-simple}\begin{aligned}
\int{\abs{\grad^{2}v}^2\abs{v}^2} \leq C\left(\widetilde{\chi},\alpha^*\right) \norm{v}_{H^{3}}^2,
\end{aligned}\end{equation}
where $C\left(\widetilde{\chi},\alpha^*\right) \rightarrow 0$ as $\alpha^*\rightarrow 0$, and,
\begin{equation}\label{Eqn-NKSobolev-full}
\begin{aligned}
&\int{\abs{\grad^{3}v}
\abs{\grad^{l_1}v}\abs{\grad^{l_2}v}\abs{\grad^{l_3}v}}\\
&+\int{\abs{\grad v}\abs{v}^2
\abs{\grad^{l_1}v}\abs{\grad^{l_2}v}\abs{\grad^{l_3}v}}\\
&+\int{\abs{\grad^{2}v}^2\left(\abs{\grad v}^2+\abs{v}^4\right)}
\end{aligned}
\leq C(\widetilde{\chi}) \frac{1}{\lambda^{7}}.
\end{equation}
\begin{comment}
the precise statements:
\[
\int{\abs{\grad^{3}v}\abs{\grad^{l_1}v}\abs{\grad^{l_2}v}\abs{\grad^{l_3}v}}
\lesssim \frac{e^{+((3)+4)\frac{m}{b(t)}}}{\lambda^{2(3)+2\delta}},
\]
\[
\int{\abs{\grad^{j_1}v}\abs{\grad^{j_2}v}\abs{\grad^{j_3}v}
\abs{\grad^{l_1}v}\abs{\grad^{l_2}v}\abs{\grad^{l_3}v}}
\lesssim \frac{e^{+(2(3)+5)\frac{m}{b(t)}}}{\lambda^{2(3)+6\delta}},
\]
\[
\int{\abs{\grad^{2}v}^2\abs{\grad v}^2}
\lesssim \frac{e^{+\frac{2}{N}((3)+2)\frac{m}{b}}}{\lambda^{2(3)}},
\]
\[
\int{\abs{\grad^{2}v}^2\abs{v}^4}
\lesssim \frac{e^{+\frac{2}{N}(1+\frac{k}{4})((3)+2)\frac{m}{b}}}{\lambda^{2(3)}}.
\]
\end{comment}
\end{lemma}
\begin{proof}
To prove (\ref{Eqn-NKSobolev-simple}), apply the three-dimensional Sobolev embeddings $H^{1} \hookrightarrow L^{6}$ and $H^{\frac{1}{2}}\hookrightarrow L^{3}$,
\[
\int{\abs{\grad^{2}v}^2\abs{v}^2} \leq
\norm{\grad^{2}v}_{L^{6}}^2\norm{v}_{L^{3}}^2.
\lesssim \norm{v}_{H^{3}}^2\norm{v}_{H^\frac{1}{2}}^2,
\]
then recall hypothesis {\bf H2.3}.
Now, consider the three LH terms of (\ref{Eqn-NKSobolev-full}) in turn, applying H\"older and three-dimensional Sobolev embeddings in each case.
\begin{enumerate}
\item
\begin{equation}\label{Proof-NKSobolev-eqn1}\begin{aligned}
\int{\abs{\grad^{3}v}\abs{\grad^{l_1}v}\abs{\grad^{l_2}v}\abs{\grad^{l_3}v}}
&\lesssim \norm{v}_{\dot{H}^{3}}\\
\prod_{j=1,2,3}\norm{\grad^{l_j}v}_{L^\frac{6}{l_j}}
&\lesssim \norm{v}_{\dot{H}^{3}}
\prod_{j=1,2,3}\norm{v}_{H^{\frac{3+l_j}{2}+\delta}},
\end{aligned}\end{equation}
where $\frac{1}{2} \gg \delta > 0$ is only necessary if $l_3 = 0$. Apply hypotheses {\bf H2.1} and {\bf H2.2},
interpolating if $\delta \neq 0$. The resulting bound is of the order $\frac{1}{\lambda^6}$.
\item
\begin{equation}\label{Proof-NKSobolev-eqn2}\begin{aligned}
\int{\abs{\grad v}\abs{v}^2
\abs{\grad^{l_1}v}\abs{\grad^{l_2}v}\abs{\grad^{l_3}v}}
&\lesssim \norm{\grad v}_{L^3}\norm{v}_{L^{18}}^2
\prod_{j=1,2,3}\norm{\grad^{l_j}v}_{L^{\frac{27}{5}\frac{1}{l_j}}}\\
&\lesssim \norm{v}_{H^\frac{3}{2}}\norm{v}_{H^\frac{4}{3}}^2
\prod_{j=1,2,3}
\norm{v}_{H^{\frac{3}{2}+\frac{4}{9}l_j+\delta}}
\end{aligned}\end{equation}
where, again, $\frac{1}{2} \gg \delta > 0$ is only necessary if $l_3 = 0$.
Apply hypotheses {\bf H2.1} through {\bf H2.3}, interpolating where necessary.
The resulting bound is of the order $\frac{1}{\lambda^\frac{8}{3}}$.
\item\begin{equation}\label{Proof-NKSobolev-eqn4}\begin{aligned}
\int{\abs{\grad^{2}v}^2\left(\abs{\grad v}^2+\abs{v}^4\right)}
&\lesssim \norm{\grad^{2}v}_{L^{6}}^2
\left(\norm{\grad v}_{L^{3}}^2
+ \norm{v}_{L^{6}}^4\right)\\
&\lesssim \norm{v}_{H^{3}}^2
\left(\norm{v}_{H^{\frac{3}{2}}}^2
+ \norm{v}_{H^1}^4\right).
\end{aligned}\end{equation}
Apply hypotheses {\bf H2.1} and {\bf H2.3}. The resulting bound is of the order $\frac{1}{\lambda^6}$.
\end{enumerate}
Finally, use hypothesis {\bf H1.3} to estimate the neglected factors of $e^\frac{1}{b}$ by a single factor of $\frac{1}{\lambda}$.
\end{proof}
Near the singular ring, and in particular on the support of $\grad\chi$, we do not have the luxury of bootstrap hypotheses. However, in this region two-dimensional type Sobolev embeddings are applicable. Coupled to the geometric decomposition, we achieve precisely the weakest usable bounds.
\begin{lemma}[Two-dimensional version of Lemma \ref{Lemma-NKSobolev}]
\label{Lemma-NSobolev}
Let $v=(1-\widetilde{\chi})u$, and suppose that $2 \geq l_1 \geq l_2 \geq l_3 \geq 0$ with $l_1+l_2+l_3 = 3$. There exists $\sigma_4 > 0$ so that,
\begin{equation}\label{Eqn-NSobolev-simple}
\int{\abs{\grad^{2}v}^2\abs{v}^2} \leq C\left(\widetilde{\chi},\widetilde{Q}_b\right)
\left(\frac{1}{\lambda^{6}} +
e^{-\frac{\sigma_4}{b}}\norm{u}_{H^{3}}^{2}\right),
\end{equation}
and,
\begin{equation}\label{Eqn-NSobolev-full}
\begin{aligned}
&\int{\abs{\grad^{3}v}\abs{\grad^{l_1}v}\abs{\grad^{l_2}v}\abs{\grad^{l_3}v}}\\
&+\int{\abs{\grad v}\abs{v}^2
\abs{\grad^{l_1}v}\abs{\grad^{l_2}v}\abs{\grad^{l_3}v}}\\
&+\int{\abs{\grad^{2}v}^2\left(\abs{\grad v}^2+\abs{v}^4\right)}\end{aligned}
\leq C\left(\widetilde{\chi},\widetilde{Q}_b\right)
\left( \frac{1}{\lambda^{8}} +
e^{-\frac{\sigma_4}{b}}\frac{1}{\lambda^2}\norm{u}_{H^{3}}^2\right).
\end{equation}
Moreover, the value of $\sigma_4>0$ is uniform over all $m>0$ sufficiently small.
\end{lemma}
\begin{proof}
Due to the concentrated support of $\Qb$ - see (\ref{Hypo1-consequenceForMu}),
\begin{equation}\label{Eqn-NSobolev-decomp}
\left(1-\widetilde{\chi}\right)u(r,z,\theta) =
\frac{1}{\lambda}
\widetilde{Q}_b\left(\frac{(r,z)-(r_0,z_0)}{\lambda}\right)e^{-i\gamma}
+ \left(1-\widetilde{\chi}\right)\widetilde{u}(r,z),
\end{equation}
which we denote by $W+w$. Due to Lemma \ref{Lemma-UnprovenProperty}, the various norms of $W$ are explicit. For example, $\norm{\grad^3 W}_{L^\infty} \leq C(\widetilde{Q}_b) \frac{1}{\lambda^{4}}$, where the constant is uniform over all $b$ sufficiently small. To prove (\ref{Eqn-NSobolev-simple}) and (\ref{Eqn-NSobolev-full}), we substitute $v = W + w$ and consider two cases: all factors are $W$, or, at least one factor is $w$. The first case is explicit and trivial. In the second case we will extract a factor that is a power of $\norm{w}_{H^1}$. Assuming $m>0$ is sufficiently small, {\bf H1.2}
will then yield the factor of $e^{-\frac{\sigma_4}{b}}$.
Throughout this proof, we preserve the correct multiplicity of $\frac{1}{\lambda}$ and $\norm{u}_{H^{3}}$ by avoiding the Sobolev embedding into $L^\infty$.
Make the substitution $v = W + w$. To prove (\ref{Eqn-NSobolev-simple}) we need to show the same bound for,
\begin{equation}\label{Proof-NSobolev-eqn1}
\int{\abs{\grad^{2}w}\abs{\grad^{2}v}\abs{v}^2}
+\int{\abs{\grad^{2}v}^2\abs{v}\abs{w}}.
\end{equation}
For the first term, apply H\"older, the two-dimensional embedding $H^\frac{1}{2} \hookrightarrow L^4$,
and interpolate,
\begin{equation}\label{Proof-NSobolev-eqn2}\begin{aligned}
\int{\abs{\grad^{2}w}\abs{\grad^{2}v}\abs{v}^2}
&\leq \norm{\grad^{2}w}_{L^4}
\norm{\grad^{2}v}_{L^4}
\norm{v}_{L^{4}}^2\\
&\lesssim \norm{w}_{H^{3-\frac{1}{2}}}
\norm{v}_{H^{3-\frac{1}{2}}}\norm{v}_{H^{\frac{1}{2}}}^2
&\lesssim
\left(\norm{w}_{H^{3}}^{\frac{3}{4}}
\norm{w}_{H^1}^{\frac{1}{4}}\right)
\norm{v}_{H^{3-\frac{1}{2}}}\norm{v}_{H^{\frac{1}{2}}}^2.
\end{aligned}
\end{equation}
Interpolate the norms in $v$ between $\norm{u}_{L^2}$ and $\norm{u}_{H^{3}}$. The factor of $\norm{w}_{H^1}$ provides a factor of $e^{-\frac{\sigma_4}{b}}$ for some $\sigma_4 > 0$.
For the second term of (\ref{Proof-NSobolev-eqn1}) follow the same strategy, except use the interpolation $\norm{w}_{H^\frac{1}{2}} \lesssim \norm{w}_{H^1}^\frac{1}{2}\norm{w}_{L^2}^{\frac{1}{2}}$. This completes the proof of (\ref{Eqn-NSobolev-simple}).
Now consider the three LH terms of (\ref{Eqn-NSobolev-full}) in turn. In each case make the substitution $v= W+ w$ and assume at least one factor is $w$.
\begin{enumerate}
\item We need to show the same bound for,
\begin{equation}\label{Proof-NSobolev-eqn3}
\int{\abs{\grad^{3}w}\abs{\grad^{l_a}W}\abs{\grad^{l_b}W}\abs{\grad^{l_c}W}}
+\int{\abs{\grad^{3}v}\abs{\grad^{l_a}v}\abs{\grad^{l_b}v}\abs{\grad^{l_c}w}},
\end{equation}
where $2\geq l_a, l_b, l_c \geq 0$, $l_a+l_b+l_c = 3$, is some permutation of $l_1, l_2, l_3$. Integrate the first term of (\ref{Proof-NSobolev-eqn3}) by parts, use H\"older and interpolate,
\[\begin{aligned}
\norm{\grad^{2}w}_{L^2}&\norm{\grad^{l_a}W\grad^{l_b}W\grad^{l_c}W}_{H^1}\\
&\lesssim
\norm{w}_{H^{3}}^\frac{1}{2}
\norm{w}_{H^1}^\frac{1}{2}
\norm{\grad^{l_a}W\grad^{l_b}W\grad^{l_c}W}_{H^1}.
\end{aligned}\]
The norms of $W$ have explicit scaling-consistent bounds of the order $\left(\frac{1}{\lambda}\right)^{(2+l_a+l_b+l_c)}$. Again, the factor $\norm{w}_{H^1}$ provides a factor of $e^{-\frac{\sigma_4}{b}}$ for some $\sigma_4 > 0$ and the resulting bound is much better than $\frac{1}{\lambda^7}$.
The remaining term of (\ref{Proof-NSobolev-eqn3}) is more difficult. Choose $p_a, p_b, p_c > 0$ such that,
\begin{equation}\label{Proof-NSobolev-eqn4}\begin{aligned}
\sum_{j=a,b,c}\frac{1}{p_j} = \frac{1}{2}
&& \text{ and }
&& \begin{aligned}
&\frac{1}{p_j} < \frac{l_j}{2} && \text{ if } l_j \neq 0,\\
&\frac{1}{p_j} < \delta_5 && \text{ if } l_j = 0,
\end{aligned}
\end{aligned}\end{equation}
where $0 <\delta_5 \ll 1$ is an arbitrary universal constant. Apply H\"older and two-dimensional Sobolev embeddings,
\begin{equation}\label{Proof-NSobolev-eqn5}\begin{aligned}
\int{\abs{\grad^{3}v}\abs{\grad^{l_a}v}\abs{\grad^{l_b}v}\abs{\grad^{l_c}w}}
&\leq
\norm{v}_{H^{3}}\norm{\grad^{l_a}v}_{L^{p_a}}\norm{\grad^{l_b}v}_{L^{p_b}}
\norm{\grad^{l_c}w}_{L^{p_c}}\\
&\lesssim
\norm{v}_{H^{3}}
\prod_{j=a,b}\norm{v}_{H^{2\left(\frac{1}{2}-\frac{1}{p_j}\right)+l_j}}\,
\norm{w}_{H^{2\left(\frac{1}{2}-\frac{1}{p_c}\right)+l_c}}.\\
\end{aligned}\end{equation}
Due to choice (\ref{Proof-NSobolev-eqn4}), the final three norms of (\ref{Proof-NSobolev-eqn5}) may be interpolated strictly between $H^{3}$ and $H^1$, or strictly between $H^1$ and $L^2$, if $l_j = 0$.
We are guaranteed a factor in $\norm{w}_{H^1}$,
\begin{equation}\label{Proof-NSobolev-Interpolate}
(\ref{Proof-NSobolev-eqn5}) \lesssim \left\{
\begin{aligned}
&\norm{u}_{H^{3}}^2\norm{u}_{H^1}^{2-C(l_c)}\norm{w}_{H^1}^{C(l_c)}
&& \text{ if } l_j \text{ all non-zero},\\
&\norm{u}_{H^{3}}^{2+C(\delta_5)}\norm{u}_{H^1}^{2-3C(\delta_5)-C(l_c)}\norm{w}_{H^1}^{C(l_c)}
&& \text{ if } l_j \text{ zero for some }j.
\end{aligned}
\right.
\end{equation}
For $m>0$ sufficiently small (relative to $\delta_5$), there is a spare factor of $e^{-\frac{\sigma_4}{b}}$, for some $\sigma_4>0$. This proves the bound for the first LH term of (\ref{Eqn-NSobolev-full}).
\item Apply H\"older and two-dimensional Sobolev embeddings, using the same values $p_j$,
\begin{equation}\begin{aligned}\label{Proof-NSobolev-eqn7}
\int{\abs{\grad v}\abs{v}^2
\abs{\grad^{l_1}v}\abs{\grad^{l_2}v}\abs{\grad^{l_3}v}}
\lesssim \norm{\grad v}_{L^4}\norm{v}_{L^8}^2
\prod_{j=1,2,3}\norm{\grad^{l_j}v}_{L^{q_j}}.
\end{aligned}\end{equation}
Recall, at least one factor of $v$ in (\ref{Proof-NSobolev-eqn7}) is infact $w$. Apply Sobolev embeddings and interpolation exactly as we did to equation (\ref{Proof-NSobolev-eqn5}).
This proves the bound for the second LH term of (\ref{Eqn-NSobolev-full}).
\item Apply H\"older and two-dimensional Sobolev,
\begin{equation}\label{Proof-NSobolev-eqn8}\begin{aligned}
\int{\abs{\grad^{2}w}\abs{\grad^{2}v}\abs{\grad v}^2}
&\leq \norm{\grad^{2}w}_{L^4}
\norm{\grad^{2}v}_{L^4}
\norm{\grad v}_{L^4}^2\\
&\lesssim \norm{w}_{H^{\frac{5}{2}}}
\norm{v}_{H^{\frac{5}{2}}}\norm{v}_{H^{\frac{3}{2}}}^2,
\end{aligned}\end{equation}
and,
\begin{equation}\label{Proof-NSobolev-eqn9}\begin{aligned}
\int{\abs{\grad^{2}w}\abs{\grad^{2}v}\abs{v}^4}
&\leq \norm{\grad^{2}w}_{L^4}
\norm{\grad^{2}v}_{L^4}
\norm{v}_{L^8}^4\\
&\lesssim \norm{w}_{H^{\frac{5}{2}}}
\norm{v}_{H^{\frac{5}{2}}}\norm{v}_{H^{\frac{3}{4}}}^4.
\end{aligned}\end{equation}
The bound for the third LH term of (\ref{Eqn-NSobolev-full}) follows from interpolation.
\end{enumerate}
\end{proof}
\begin{lemma}[$H^{3}$ Energy Identity]\label{Lemma-HnkEnergy}
Denote the third-order energy by,
\begin{equation}\label{DefnEqn-Enk}
E_{3}(u) =
\int{\abs{\grad^{3}u}^2}
-\left(2\int{\abs{\grad^{2}u}^2\abs{u}^2}
+Re\int{(\grad^{2}\overline{u})^2u^2}\right).
\end{equation}
Then,
\begin{equation}\label{Eqn-HnkEnergy}\begin{aligned}
\frac{1}{C}\abs{\frac{d}{dt}E_{3}(u)}
\leq &
\int{\abs{\grad^{3}u}
\abs{\grad^{l_1}u}\abs{\grad^{l_2}u}\abs{\grad^{l_3}u}}\\
&+\int{\left(\abs{\grad u}\abs{u}^2\right)
\abs{\grad^{l_1}u}\abs{\grad^{l_2}u}\abs{\grad^{l_3}u}}\\
&+ \int{\abs{\grad^{2}u}^2\left(\abs{\grad u}^2+\abs{u}^4\right)},
\end{aligned}
\end{equation}
where the right side is implicitly summed over $2 \geq l_1 \geq l_2 \geq l_3 \geq 0$ with $l_1+l_2+l_3 = 3$.
\end{lemma}
\begin{proof}
We refer to the RHS of (\ref{Eqn-HnkEnergy}) as error terms of type I, II, and III respectively.
By direct calculation,
\begin{equation}\label{Proof-HnkEnergy-eqn1}
\begin{aligned}
\frac{1}{2}\frac{d}{dt}\left(\int{\abs{\grad^{3}u}^2}\right)
& = -Im
\int{\grad^{3}\left(\laplacian u
+u\abs{u}^2\right)\,\grad^{3}\overline{u}}\\
& = -2\,Im\int{\grad\left(\grad^{2}u\,\abs{u}^2\right)\,\grad^{3}\overline{u}}\\
&\mspace{36.0mu}-Im\int{\grad\left(\grad^{2}\overline{u}u^2\right)\,\grad^{3}\overline{u}}\\
&\mspace{54.0mu}+ \text{terms of the form }
\int{\grad \left(\grad u \grad u\,u\right)\grad^{3}\overline{u}}.
\end{aligned}\end{equation}
The final terms of (\ref{Proof-HnkEnergy-eqn1}) are error of type I. Regarding the first RH term of (\ref{Proof-HnkEnergy-eqn1}),
\begin{equation}\label{Proof-HnkEnergy-eqn2}\begin{aligned}
-2\,Im\int{\grad\left(\grad^{2}u\,\abs{u}^2\right)\,\grad^{3}\overline{u}}
\;=\; & 2\,Im\int{\grad^{2}u\,\abs{u}^2\grad^{2}\laplacian \overline{u}}\\
=\; & \int{\frac{d}{dt}\left(\abs{\grad^{2}u}^2\right)\,\abs{u}^2}
-2\,Im\int{\grad^{2}u\,\abs{u}^2\grad^{2}\left(\overline{u}\abs{u}^2\right)}.
\end{aligned}
\end{equation}
Recognize the last term of (\ref{Proof-HnkEnergy-eqn2}) as error of type II and III. Regarding the other term,
\begin{equation}\label{Proof-HnkEnergy-eqn3}\begin{aligned}
\int{\frac{d}{dt}\left(\abs{\grad^{2}u}^2\right)\,\abs{u}^2}
=&\frac{d}{dt}\left(\int{\abs{\grad^{2}u}^2\,\abs{u}^2}\right)
+2\,Im\int{\abs{\grad^{2}u}^2\left(\laplacian u + u\abs{u}^2\right) \overline{u}}.
\end{aligned}\end{equation}
After integration by parts, we recognize the final term of (\ref{Proof-HnkEnergy-eqn3}) as error of type I and III. We have shown that,
\[
-2\,Im\int{\grad\left(\grad^{2}u\,\abs{u}^2\right)\,\grad^{3}\overline{u}}
= \frac{1}{2}\frac{d}{dt}\left(\int{\abs{\grad^{2}u}^2\,\abs{u}^2}\right),
\]
up to error terms. It is virtually the same calculation to show that,
\[
-Im\int{\grad\left(\grad^{2}\overline{u}u^2\right)\,\grad^{3}\overline{u}}
= \frac{1}{2}\frac{d}{dt}\left( Re\int{(\grad^{2}\overline{u})^2u^2} \right),
\]
also up to error terms of type I, II, and III. This completes the proof of (\ref{Eqn-HnkEnergy}).
\end{proof}
Now we simply combine the previous three Lemmas.
Equations (\ref{Eqn-NKSobolev-simple}) and (\ref{Eqn-NSobolev-simple}) prove that $E_3 \approx \norm{u}_{H^3}$. Equations (\ref{Eqn-NKSobolev-full}) and (\ref{Eqn-NSobolev-full}) control $\frac{d}{dt}E_3$. Integrate the bound on $\frac{d}{dt}E_3$ using Lemma \ref{Lemma-lambdaIntegral}, with $\sigma_2 < \min(\sigma_4, 2m)$. Choose $m'>0$ to be any value, $m-\frac{\sigma_4}{2}<m'<m$. Assuming $\alpha^*$ is sufficiently small (depending on the choice of $m'$), we have proven,
\begin{lemma}[Controlled Growth of $H^{3}$]\label{Lemma-ControlledHnk}
For all $t\in[0,T_{hyp})$,
\begin{equation}\label{Eqn-Hnk}
\norm{u(t)}_{H^{3}(\real^{3})}
< \frac{e^{\frac{m'}{b(t)}}}{\lambda^{3}(t)}.
\end{equation}
That is, statement {{\bf I2.1}}.
\end{lemma}
\begin{remark}[Higher Dimensions]\label{Remark-HigherDim-NthEnergy}
For the higher dimensional case, define,
\[
E_N(u) = \int{\abs{\grad^Nu}^2} - \left(2\int{\abs{\grad^{N-1}u}^2\abs{u}^2} + Re\int{\left(\grad^{N-1}u\right)^2\left(\overline{u}\right)^2}\right).
\]
Then $\abs{\frac{d}{dt}E_N(u)}$ may be bounded in the same way by generalizing error of type II to include,
\[\begin{aligned}
\int{\left(\abs{\grad^{k_1}u}\abs{\grad^{k_2}u}\abs{\grad^{k_3}u}\right)
\abs{\grad^{l_1}u}\abs{\grad^{l_2}u}\abs{\grad^{l_3}}}
&&\text{ with }&&k_1+k_2+k_3 = N-2.
\end{aligned}\]
For the calculation with quintic nonlinearity, see \cite[Lemma 1]{RaphaelSzeftel-StandingRingNDimQuintic-08}.
\end{remark}
\begin{comment}
\begin{remark}[Constraints on the form of Hypothesis (\ref{Hypo2-Hnk})]
\label{Remark-NoChoicePowerOfHnkLambda}
The choice of bootstrap hypothesis (\ref{Hypo2-Hnk}) is no accident - it is necessary due to our estimate (\ref{Eqn-NSobolev-full}) for the 'type I' error of (\ref{Eqn-HnkEnergy}). The alternate hypothesis:
\[
\norm{u}_{H^{3}} < C\frac{1}{\lambda^{3}},
\]
would suffer a loss during time-integration. Likewise, if we choose the alternate hypothesis:
\begin{equation}\label{Remark-NoChoicePowerOfHnkLambda-eqn1}
\norm{u}_{H^{3}} < C\frac{1}{\lambda^{3+\delta_0}},
\end{equation}
then, from (\ref{Proof-NSobolev-Interpolate}) when $l_k = 0$, the resulting form of (\ref{Eqn-NSobolev-full}) has too high a power of $\frac{1}{\lambda}$. It is possible in some cases to algebraically cancel terms with $l_k = 0$ during the calculation of (\ref{Eqn-HnkEnergy}). However, there is nothing to gain, as an argument in $v_W+w$ is still necessary to prove (\ref{Eqn-NSobolev-full}).
\end{remark}
\end{comment}
\subsection{Behaviour away from both Infinity and the Singularity}
\label{Subsection-Annular}
This section we concentrate on the interface between the singular ring and the truly three-dimensional region that contains the origin. On this interface, away from $r=0$, the dynamics remains essentially two-dimensional and $L^2$-critical
\begin{lemma}[Two-dimensional endpoint Sobolev control away from the singularity]
\label{Lemma-NdimEndpointSobolev}
For $\sigma_5 > 0$, and a smooth cutoff function $\varphi$ compactly supported away from both the singular ring and the origin,
\begin{equation}\label{Eqn-NdimEndpointSobolev}
\norm{\varphi u(t)}_{L^\infty(\real^3)}
\leq C(\sigma_5, \varphi)
e^{+\frac{\sigma_5}{b(t)}}
\left(\norm{\widetilde{u}(t)}_{H^1(\real^3)}
+ \frac{\Gamma^\frac{1}{2}_{b(t)}}{\lambda(t)}\right).
\end{equation}
This is a two-dimensional type of estimate due to the support of $\varphi$.
\end{lemma}
The key feature of Lemma \ref{Lemma-NdimEndpointSobolev} is that we may avoid the Sobolev embedding $H^{1+\epsilon}(\real^2) \hookrightarrow L^\infty(\real^2)$. At the order of the blowup parameter $\lambda$, equation (\ref{Eqn-NdimEndpointSobolev}) is consistent with scaling. In the case of radial symmetry, such as \cite{R06,RaphaelSzeftel-StandingRingNDimQuintic-08}, Strauss' radial embedding is used instead.
\begin{proof}[Proof of Lemma \ref{Lemma-NdimEndpointSobolev}]
We adapt an argument of Brezis \& Gallou\"et.
Our estimate is for a fixed time $t\in[0,T_{hyp})$.
Choose $R = \norm{\widetilde{u}(t)}_{H^1} + \frac{\Gamma^\frac{1}{2}_{b(t)}}{\lambda(t)} \gg 0$. Denote $v = \varphi u$ and partition phase space,
\[
\abs{v} \leq \norm{\hat{v}}_{L^1} =
\int_{\abs{\xi} \leq R}{\abs{\hat{v}(\xi)}\,d\xi}
+\int_{\abs{\xi} > R}{\abs{\hat{v}(\xi)}\,d\xi}
\]
Rewrite the low frequencies and apply H\"older,
\begin{equation}\label{Eqn-NdimEndpoint-eqn1}\begin{aligned}
\int_{\abs{\xi} \leq R}{\abs{\hat{v}}\,d\xi}
&= \int_{\abs{\xi} \leq R}{
\left({\langle\xi\rangle}^{\frac{1}{2}}
\abs{\hat{v}}^{\frac{1}{2}}\right)
\left(\abs{\hat{v}}^{\frac{1}{2}}\right)
\left(\frac{1}{{\langle\xi\rangle}^{\frac{1}{2}}}\right)\,d\xi
}\\
&\leq \norm{v}_{H^{1}}^{\frac{1}{2}}
\norm{v}_{L^2}^{\frac{1}{2}}
\left(\int_{\abs{\xi} \leq R}{\frac{1}{{\langle\xi\rangle}}\,d\xi}
\right)^\frac{1}{2},
\end{aligned}\end{equation}
where ${\langle\xi\rangle}$ denotes $\sqrt{1 + \abs{\xi}^2}$. Note the final integral of (\ref{Eqn-NdimEndpoint-eqn1}) is,
$
\int_{\abs{\xi} \leq R}{\frac{1}{{\langle\xi\rangle}}\,d\xi}
\leq \int_0^R{\frac{1}{\rho}\rho\,d\rho}
= R.
$
Apply a similar argument to high frequencies, with parameter $\nu(\sigma_5,m)>1$ to be determined,
\begin{equation}\label{Eqn-NdimEndpoint-eqn3}\begin{aligned}
\int_{\abs{\xi} > R}{\abs{\hat{v}}\,d\xi}
&= \int_{\abs{\xi} > R}{\left({\langle\xi\rangle}^\nu\abs{\hat{v}}\right)
\frac{1}{{\langle\xi\rangle}^\nu}\,d\xi}\\
&\lesssim \norm{v}_{H^\nu}\,
\left(\int_R^{+\infty}{\frac{1}{{\langle \rho\rangle}^{2\nu}}\rho\,d\rho}\right)^\frac{1}{2}\\
&\leq \frac{1}{2(\nu-1)}\norm{v}_{H^\nu}\,\frac{1}{R^{\nu-1}}\\
&\lesssim \frac{1}{2(\nu-1)}
\left(\norm{v}_{H^1}^{2-\nu}R^{\nu - 1}\right)
\left(\frac{\norm{v}_{H^2}^{\nu-1}}{R^{2(\nu-1)}}\right).
\end{aligned}\end{equation}
Due to hypothesis {\bf H2.1}
and $\Gamma_b$-estimate (\ref{Eqn-GammaBEstimate}), the final term of (\ref{Eqn-NdimEndpoint-eqn3}) is bounded by $e^{+\frac{\sigma_5}{b(t)}}$ for any choice of $\nu > 1$ sufficiently small.
\end{proof}
\begin{definition}[Cutoffs to cover $Supp\left(\grad\chi\right)$]\label{Defn-AnnularCutoffs}
Fix seven smooth cylindrically symmetric cutoff functions, $\psi^{(0)}$, $\psi^{(\frac{1}{2})}$, $\psi^{(1)}$, $\varphi^{(\frac{5}{2})}$, $\varphi^{(2)}$, $\varphi^{(\frac{3}{2})}$, $\varphi^{(1)}$, such that,
\begin{enumerate}
\item {\it They cover the support of $\grad\chi$.}
Each function is $1$ on, $\left\{ \frac{1}{3} < \abs{(r,z)-(1,0)} < \frac{2}{3}\right\}$.
\item {\it Tails do not overlap.}
The support of each cutoff is contained where the previous cutoff is $1$.
\item {\it Supported away from both the singularity and the origin.}
The largest support, that of $\psi^{(0)}$, is contained in, $\left\{ \frac{1}{7} < \abs{(r,z) - (1,0)} < \frac{6}{7}\right\}$.
\end{enumerate}
\end{definition}
\begin{lemma}[Annular $H^\frac{1}{2}$ control - the crucial first step]
\label{Lemma-Annular12}
For all $t\in[0,T_{hyp})$,
\begin{equation}\label{Eqn-Annular12}
\norm{\psi^{(\frac{1}{2})}u}_{H^\frac{1}{2}} \lesssim \frac{1}{\lambda^{C(\alpha^*)}(t)},
\end{equation}
where $C(\alpha^*) \rightarrow 0$ as $\alpha^* \rightarrow 0$.
\end{lemma}
This is the first proof that any behaviour better than scaling extends beyond the support of hypotheses {\bf H2.2} and {\bf H2.3}.
\begin{remark}[Analogue in \cite{R06,RaphaelSzeftel-StandingRingNDimQuintic-08}]
In radial cases, one proves Lemma \ref{Lemma-Annular12} for $H^\nu$, $\nu < \frac{1}{2}$. The subsequent $H^\frac{1}{2}$ bound, for example \cite[Lemma 10]{R06}, should be seen as comparable to forthcoming Lemma \ref{Lemma-AnnularN2}.
\end{remark}
\begin{proof}[Proof of Lemma \ref{Lemma-Annular12}]
By direct calculation,
\begin{equation}\label{Proof-Annular12-Gronwall}
\frac{1}{2}\frac{d}{dt}\norm{\psi^{(\frac{1}{2})}u}_{\dot{H}^\frac{1}{2}}^2 =
Im\left(\int{D^\frac{1}{2}\left(u\laplacian\psi^{(\frac{1}{2})}
+ 2\grad\psi^{(\frac{1}{2})}\cdot\grad u
- \psi^{(\frac{1}{2})}u\,\abs{u}^2\right)
\,D^\frac{1}{2}\left(\psi^{(\frac{1}{2})}\overline{u}\right)}\right).
\end{equation}
Estimate the first and second RH terms of (\ref{Proof-Annular12-Gronwall}),
\begin{align}
\notag \norm{D^\frac{1}{2}\left(u\laplacian\psi^{(\frac{1}{2})}\right)}_{L^2}
\norm{\psi^{(\frac{1}{2})}u}_{H^\frac{1}{2}}
&\leq C\left(\psi^{(\frac{1}{2})}\right)\norm{\psi^{(0)}u}_{H^\frac{1}{2}}
\norm{\psi^{(\frac{1}{2})}u}_{H^\frac{1}{2}},\\
\notag \norm{\grad\psi^{(\frac{1}{2})}\cdot\grad u}_{L^2}
\norm{\psi^{(\frac{1}{2})}u}_{H^1}
&\leq C\left(\psi^{(\frac{1}{2})}\right)\norm{\psi^{(0)}u}_{H^{1}}^2.
\end{align}
The nonlinear term of (\ref{Proof-Annular12-Gronwall}) does not enjoy any real-valued cancellations, as the operator $D$ does not have an exact Leibniz property.
Apply standard commutation estimates,
\begin{equation}\label{Proof-Annular12-eqnNeedBrezis}\begin{aligned}
\norm{D^\frac{1}{2}\left(\psi^{(\frac{1}{2})}u\,\abs{u}^2\right)}_{L^2}
\lesssim
&
\norm{\psi^{(\frac{1}{2})}u}_{H^\frac{1}{2}}\norm{\psi^{(0)}u}_{L^\infty(\real^2)}^2
+\norm{\psi^{(\frac{1}{2})}u}_{L^4}
\norm{\psi^{(0)}{u}}_{W^{\frac{1}{2},4}}\norm{\psi^{(0)}u}_{L^\infty(\real^2)}\\
\lesssim &
\norm{\psi^{(\frac{1}{2})}u}_{H^{\frac{1}{2}}}\left(
\norm{\psi^{(0)}u}_{L^\infty(\real^2)}^2
+ \norm{\psi^{(0)}u}_{H^1}\norm{\psi^{(0)}u}_{L^\infty(\real^2)}
\right).
\end{aligned}\end{equation}
From support away from the singularity, $\psi^{(0)}u = \psi^{(0)}\widetilde{u}$, and we may apply the endpoint estimate of Lemma \ref{Lemma-NdimEndpointSobolev}.
Denote $\norm{\psi^{(\frac{1}{2})}u(t)}_{\dot{H}^\frac{1}{2}}$ by $f$. We have the simple ODE,
\begin{equation}\label{Proof-Annular12-simpleODE}\begin{aligned}
\frac{1}{2}\frac{d}{dt}\left(f^2\right) \leq
&C\left(\psi^{(\frac{1}{2})}\right)\left(
f\,\norm{\widetilde{u}(t)}_{H^1}^\frac{1}{2}
\norm{\widetilde{u}(t)}_{L^2}^\frac{1}{2}
+ \norm{\widetilde{u}(t)}_{H^1}^2
\right)\\
&+C\left(\sigma_5,\psi^{(0)}\right)f^2\,
e^{+\frac{\sigma_5}{b(t)}}
\left(\norm{\widetilde{u}(t)}_{H^1}^2
+ \frac{\Gamma_{b(t)}}{\lambda^2(t)}\right).
\end{aligned}\end{equation}
The final term is dominant. After integration by Corollary \ref{Corollary-lambdaIntegral2version2},
\[
\norm{\psi^{(\frac{1}{2})}u(t)}_{{H}^\frac{1}{2}}
\lesssim e^{\left\lbrack C\left(\alpha^*\right)\,C\left(\sigma_5,\psi^{(0)}\right)e^{+\frac{\sigma_5}{b(t)}}\right\rbrack}.
\]
To complete the proof, choose $\sigma_5 = \frac{\pi}{10}$ and recall the log-log rate {\bf H1.3}.
\end{proof}
\begin{remark}[Justification for Lemma \ref{Lemma-NdimEndpointSobolev}]
\label{Remark-NormalBrezisFails}
The open nature of hypothesis {\bf H1.3} is an essential feature of any modulation argument. It is for this reason that we must be free to choose $\sigma_5$. The standard Brezis-Gallou\"et estimate,
$
\norm{v}_{L^\infty(\real^2)} \lesssim \norm{v}_{H^1}\sqrt{\log\left(\norm{v}_{H^2}\right)},
$ would not suffice to prove Lemma \ref{Lemma-Annular12}.
\end{remark}
We now reformulate the calculation of equation (\ref{Proof-Annular12-Gronwall}) for repeated application.
\begin{lemma}[Standard Gronwall Argument]\label{Lemma-GronwallArg}
Let $\psi^A$ be supported where $\psi^B\equiv 1$, let $I$ be any subinterval of $[0,T_{hyp})$, and let $\nu\geq 0$. Then,
\begin{equation}\label{Eqn-GronwallTool}
\norm{\psi^Au}_{L^\infty_IH^\nu} \leq C(\psi^A)\left(
\norm{\psi^Bu_0}_{H^\nu} + \abs{I} +
\norm{\psi^Bu}_{L^2_IH^{\nu+\frac{1}{2}}} +
\norm{\psi^Au\,\abs{u}^2}_{L^1_IH^{\nu}}\right).
\end{equation}
\end{lemma}
\begin{comment}
\begin{proof}
Denote $\psi^Au$ by $v$.
Then by direct calculation,
\begin{equation}\label{Proof-Gronwall-eqn2}
\frac{1}{2}\frac{d}{dt}\norm{D^{\nu}v}_{L^2}^2 =
Im\left(\int{D^{\nu}\left(u\laplacian\psi^A + 2\grad\psi^A\cdot\grad u - v\,\abs{u}^2\right)\,D^{\nu}\overline{v}}\right).
\end{equation}
Estimate the three terms of equation (\ref{Proof-Gronwall-eqn2}) respectively by,
\[
\norm{D^{\nu}(u\laplacian\psi^A)}_{L^2}\norm{v}_{H^\nu}
\leq C(\psi^A)\norm{\psi^Bu}_{H^\nu}\norm{v}_{H^\nu},\]
\[
\norm{D^{\nu-\frac{1}{2}}\left(\grad\psi^A\cdot\grad u\right)}_{L^2}\norm{v}_{H^{\nu+\frac{1}{2}}}
\leq C(\psi^A)\norm{\psi^Bu}_{H^{\nu+\frac{1}{2}}}^2,\]
and $\norm{D^{\nu}\left(v\,\abs{u}^2\right)}_{L^2}\norm{v}_{H^\nu}$.
Integration in time gives (\ref{Eqn-GronwallTool}).
\end{proof}
\end{comment}
\begin{lemma}[Annular $H^1$ control - propagation of Lemma \ref{Lemma-Annular12}]
\label{Lemma-AnnularN2}
There exists $\sigma_6 > 0$, universal for all $m>0$ sufficiently small, such that for all $t\in[0,T_{hyp})$,
\begin{equation}\label{Eqn-AnnularN2}
\norm{\psi^{(1)}u(t)}_{H^1} < C(\alpha^*)\frac{e^{-\frac{\sigma_6}{b(t)}}}{\lambda^\frac{1}{2}(t)},
\end{equation}
where $C(\alpha^*) \rightarrow 0$ as $\alpha^* \rightarrow 0$.
\end{lemma}
\begin{proof}
Apply the standard Gronwall argument, equation (\ref{Eqn-GronwallTool}), for $\nu=1$, $I = [0,t<T_{hyp}]$, $\psi^A = \psi^{(1)}$, and $\psi^B = \psi^{(\frac{1}{2})}$. Note that $\psi^{(1)}u = \psi^{(1)}\tilde{u}$. Through interpolation and hypotheses {\bf H1.2} and {\bf H2.1},
\begin{equation}\label{Proof-AnnularN2-eqn1}\begin{aligned}
\norm{\psi^{(1)}u}_{L^2_IH^{1+\frac{1}{2}}}
\lesssim& \left(\int{
\norm{\tilde{u}}_{H^1}^{2-\frac{1}{2}}
\norm{\tilde{u}}_{H^{3}}^{\frac{1}{2}}
}\right)^\frac{1}{2}
&\lesssim& \left(\int{
e^{-\left(\frac{1}{4}-\frac{m}{2}\right)\frac{1}{b}}\frac{1}{\lambda^3}
}\right)^\frac{1}{2}.
\end{aligned}\end{equation}
Assuming $m>0$ is sufficiently small, apply integrability Lemma \ref{Lemma-lambdaIntegral} for $\sigma_2 > 0$, also sufficiently small.
Regarding the final term of (\ref{Eqn-GronwallTool}), apply H\"older, two-dimensional Sobolev embedding, and interpolate,
\begin{equation}\label{Proof-AnnularN2-eqn2}\begin{aligned}
\norm{\psi^{(1)}u\,\abs{u}^2}_{H^1} \lesssim &
\norm{\grad\left(\psi^{(\frac{1}{2})}u\right)\,\left(\psi^{(\frac{1}{2})}u\right)^2}_{L^2}\\
\lesssim &
\norm{\grad\left(\psi^{(\frac{1}{2})}u\right)}_{L^4}\norm{\psi^{(\frac{1}{2})}u}_{L^8}^2\\
\lesssim &
\norm{\psi^{(\frac{1}{2})}u}_{H^\frac{3}{2}}\norm{\psi^{(\frac{1}{2})}u}_{H^\frac{3}{4}}^2\\
\lesssim &
\norm{u}_{H^{3}}^{\frac{3}{5}}
\norm{\psi^{(\frac{1}{2})}u}_{H^\frac{1}{2}}^{3-\frac{3}{5}}
& \lesssim\frac{1}{\lambda^{\frac{9}{5}+C(\alpha^*)}(t)},
\end{aligned}\end{equation}
where the final inequality is due to hypothesis {\bf H2.1} and Lemma \ref{Lemma-Annular12}. Apply Lemma \ref{Lemma-lambdaIntegralCrude}.
\end{proof}
\begin{comment}
\begin{remark}
One can iterate the proof of Lemma \ref{Lemma-AnnularN2} to prove the $\frac{1}{2}$-derivative improvement propagates up to $\norm{\psi^{(\frac{5}{2})}u}_{H^{\frac{5}{2}}}$, with appropriately shrinking cutoff functions. Instead, we will introduce new cutoff functions and iterate in the opposite direction. The bound of Lemma \ref{Lemma-AnnularN2} (compared with Lemma \ref{Lemma-Annular12}) will allow us to run these new iterates with the correct 'step-size'.
\end{remark}
\end{comment}
\begin{remark}[Scheme for the remainder of Chapter \ref{Section-BootAtInfty}]
The proof of Lemma \ref{Lemma-AnnularN2} may be repeated, with a shrunken cutoff and $H^\frac{3}{2}$ in place of $H^1$. However, due to the new version of equation (\ref{Proof-AnnularN2-eqn1}), iteration to higher norms will not yield more than the same $\frac{1}{2}$-derivative improvement over scaling.
Instead, we switch direction. Starting with {\bf I2.1}, at each stage the previous iterate will give progressively better control on the equivalent of (\ref{Proof-AnnularN2-eqn1}). Lemma \ref{Lemma-AnnularN2} will be used to help control the equivalent of equation (\ref{Proof-AnnularN2-eqn2}).
\begin{comment}
We will apply Lemma \ref{Lemma-GronwallArg} iteratively, at each stage choosing both a different norm $H^s$ and a new cutoff $\psi^A$ with smaller support. To begin we will 'work up', starting with $H^\frac{1}{2}$, in which cases $\norm{\psi^Bu}_{L^2_IH^{s+\frac{1}{2}}}$ will limit the obtainable bounds. In particular, we may hope for approximately $\frac{1}{2}$-derivative improvement on the bound for $H^1$, where here $1$ should be seen as half the $L^2$-critical dimension.
Then, starting with $H^{\frac{5}{2}}$ and 'working down', the main obstruction is the nonlinear term $\norm{v\,\abs{u}^2}_{L^1_IH^s}$, for which we will apply the improved bound on $H^1$ previously obtained.
\end{comment}
\end{remark}
\begin{lemma}[Moser-type Product Estimate]\label{Lemma-MoserType}
Let $v \in H^{\nu+\frac{1}{2}}(\real^d)$ for some $\nu \geq \frac{d-1}{2}$, not necessarily an integer. Then,
\begin{equation}\label{Eqn-MoserEst}
\norm{v^3}_{H^\nu} \lesssim
\norm{v}_{H^{\nu+\frac{1}{2}}}
\norm{v}_{H^{\frac{d}{2}}}^2.
\end{equation}
\end{lemma}
\begin{comment}
\begin{proof}[Partial Proof of Lemma \ref{Lemma-MoserType}]
For $\nu \geq \frac{d}{2}$, (\ref{Eqn-MoserEst}) follows from,
\begin{equation}\label{Proof-MoserEst-pre-eqn1}
\norm{v^3}_{H^\nu} \lesssim \norm{v}_{H^\nu}\norm{v}_{L^\infty}^2,
\end{equation}
the Sobolev embedding $H^{\frac{d}{2}+\delta} \hookrightarrow L^\infty$ for some small $\delta$ and then interpolation. See \cite[page 84]{AlinhacGerard} for a Littlewood-Paley proof of (\ref{Proof-MoserEst-pre-eqn1}),
or later in the same volume for the Nash-Moser theorem.
The case $\nu = \frac{d-1}{2}$ we need to be a little more careful. We will use this case only for $d=3$ in the proof of Lemma \ref{Lemma-OriginHLower}.
For completeness, a proof of Lemma \ref{Lemma-MoserType} based on H\"older and Sobolev and including the case $\nu = \frac{d-1}{2}$ is provided in Appendix \ref{Appendix-AuxilliaryProof}.
\end{proof}
\end{comment}
\begin{lemma}[{\bf I2.2} and {\bf I2.3} on the support of $\grad \chi$]
\label{Lemma-AnnularHlower}
For all $t \in [0,T_{hyp})$:
\begin{equation}\label{Eqn-AnnularHlower}
\norm{\varphi^{(3-\kappa)}u}_{H^{3-\kappa}} < C(\alpha^*) \frac{e^{(1+\kappa)\frac{m'}{b(t)}}}{\lambda^{3-2\kappa}},
\end{equation}
for each half integer $\frac{1}{2} \leq \kappa < \frac{3}{2}$,
\begin{equation}\label{Eqn-AnnularHlowerN2}
\norm{\varphi^{(\frac{3}{2})} u(t)}_{H^\frac{3}{2}} < C(\alpha^*) e^{+\frac{2m' + \pi}{b(t)}},
\end{equation}
and,
\begin{equation}\label{Eqn-AnnularHCrit}
\norm{\varphi^{(1)}u}_{H^{1}} < C(\alpha^*) \left(\alpha^*\right)^{\frac{1}{5}},
\end{equation}
where in each case $C(\alpha^*) \rightarrow 0$ as $\alpha^* \rightarrow 0$.
\end{lemma}
\begin{proof}
We prove (\ref{Eqn-AnnularHlower}) by induction in $\kappa$. The base case $\kappa = 0$ is Lemma \ref{Lemma-ControlledHnk}.
Hypothesize that (\ref{Eqn-AnnularHlower}) holds for $\kappa - \frac{1}{2}$, some $\kappa \geq \frac{1}{2}$. Denote $\nu = 3-\kappa$, and apply the standard Gronwall argument for $I=[0,t<T_{hyp}]$, $\psi^A =\phi^{(\nu)}$ and $\psi^B = \phi^{(\nu+\frac{1}{2})}$,
\begin{equation}\label{Proof-AnnularLower-eqn1}
\norm{\varphi^{(\nu)}u}_{H^\nu} \lesssim
\norm{\chi_0u_0}_{H^\nu}
+ \norm{\varphi^{(\nu+\frac{1}{2})}u}_{L^2_tH^{\nu+\frac{1}{2}}}
+ \norm{\varphi^{(\nu)}u\,\abs{u}^2}_{L^1_tH^{\nu}}.
\end{equation}
Apply our induction hypothesis to the second RH term of (\ref{Proof-AnnularLower-eqn1}),
\begin{equation}\label{Proof-AnnularLower-eqn1.2}
\norm{\varphi^{(\nu+\frac{1}{2})}u}_{L^2_tH^{\nu+\frac{1}{2}}}
\lesssim \left(\int_I{\left(\frac{e^{\left(1+(\kappa-\frac{1}{2})\right)\frac{m'}{b(\tau)}}}{\lambda^{3-2(\kappa-\frac{1}{2})}(\tau)}\right)^2d\,\tau}\right)^\frac{1}{2}
\lesssim
\left(\frac{e^{(1+\kappa)\frac{m'}{b(t)}}}{\lambda^{3-2\kappa}(t)}\right)
\left(e^{\frac{\sigma_2-m'}{b(t)}}\right)^\frac{1}{2}.
\end{equation}
where, since $\kappa < \frac{3}{2}$, we applied Lemma \ref{Lemma-lambdaIntegral} for some $\sigma_2 < m'$.
Examine the final term of (\ref{Proof-AnnularLower-eqn1}).
Note that, $\varphi^{(\nu)}u = \varphi^{(\nu)}\left( \varphi^{(\nu+\frac{1}{2})}u \right)$.
Apply the Moser-type estimate of Lemma \ref{Lemma-MoserType}
and inject both the $H^1$ control of Lemma \ref{Lemma-AnnularN2} and the induction hypothesis,
\begin{equation}\label{Proof-AnnularLower-useMoser}\begin{aligned}
\norm{\varphi^{(\nu)}u\,\abs{u}^2}_{H^\nu} \lesssim &
\norm{\varphi^{(\nu+\frac{1}{2})}u}_{H^{\nu+\frac{1}{2}}}
\norm{\varphi^{(\nu+\frac{1}{2})}u}_{H^1}^2\\
\lesssim &
\frac{e^{+\frac{(1+(\kappa-\frac{1}{2}))m'}{b}}}{\lambda^{3-2(\kappa-\frac{1}{2})}}
\frac{e^{-\frac{2\sigma_6}{b}}}{\lambda}
& = \frac{e^{\frac{(\text{neg})}{b}}}{\lambda^2}\frac{1}{\lambda^{3-2\kappa}},
\end{aligned}\end{equation}
where we made the assumption $m>0$ is sufficiently small relative to $\sigma_6$.
Finally, apply Lemma \ref{Lemma-lambdaIntegral} for some $\sigma_2$ less than the negative exponent. This completes the proof of (\ref{Eqn-AnnularHlower}).
To prove (\ref{Eqn-AnnularHlowerN2}) let $\kappa = \frac{3}{2}$. We proceed exactly as above, using (\ref{Eqn-AnnularHlower}) in place of the induction hypothesis, and applying Corollary \ref{Corollary-lambdaIntegral2version2} in place of Lemma \ref{Lemma-lambdaIntegral}.
To prove (\ref{Eqn-AnnularHCrit}), let $\kappa = 2$. We proceed exactly as above using (\ref{Eqn-AnnularHlowerN2}) in place of the induction hypothesis, and applying Lemma \ref{Lemma-lambdaIntegralCrude} in place of Lemma \ref{Lemma-lambdaIntegral}.
\end{proof}
\subsection{Improved Behaviour at Infinity}
\label{Subsection-Infty}
With Lemma \ref{Lemma-AnnularHlower} covering the support of $\grad\chi$, we prove the corresponding result for $\chi$ by similar methods. Note the argument is now in three-dimensions.
\begin{proof}[Proof of {\bf I2.2} and {\bf I2.3}]
We revisit the proof of the standard Gronwall argument.
Let $I = [0,t<T_{hyp}]$, $\nu \geq 0$, and denote $v = \chi u$. With equation (\ref{Eqn-NLS}),
\begin{equation}\label{Proof-OriginHLower-eqn1}
iv_t+\laplacian v +v\abs{v}^2= u\laplacian\chi + 2\grad\chi\cdot\grad u + \left(\chi^2 - 1\right)\chi u\abs{u}^2.
\end{equation}
Note that the terms on the RHS of (\ref{Proof-OriginHLower-eqn1}) are localized to the support of $\grad\chi$, a region of two-dimensional character where $\varphi^{(1)} \equiv 1$.
By direct calculation,
\begin{equation}\label{Proof-OriginHLower-miniGron}\begin{aligned}
\frac{1}{2}\norm{\chi u}_{L^\infty_IH^\nu} \leq
&\norm{\chi u_0}_{H^\nu} + \norm{\chi u\abs{\chi u}^2}_{L^1_IH^\nu}\\
&+
C(\chi)\left(
\norm{\varphi^{(1)}u_0}_{H^\nu} + \abs{I}
+\norm{\varphi^{(1)}u}_{L^2_IH^{\nu+\frac{1}{2}}}
+\norm{\varphi^{(1)}u\abs{\varphi^{(1)}u}^2}_{L^1_IH^\nu}
\right).
\end{aligned}\end{equation}
Consider $\nu = 3-\kappa$ for some $\frac{1}{2}\leq\kappa\leq 2$.
Due to Definition \ref{Defn-AnnularCutoffs}, all the conclusions of Lemma \ref{Lemma-AnnularHlower} apply to $\varphi^{(1)}u$, which we use in place of an induction hypothesis to control the second line of (\ref{Proof-OriginHLower-miniGron}) exactly as we did equation (\ref{Proof-AnnularLower-eqn1}). These terms will give the largest contribution.
Finally, examine the term nonlinear in $\chi u$.
Apply the Moser-type estimate of Lemma \ref{Lemma-MoserType},
interpolate, and inject hypotheses {\bf H2.2},
\begin{equation}\label{Proof-OriginHLower-useMoser}\begin{aligned}
\norm{\chi u\,\abs{\chi u}^2}_{L^1_IH^\nu}
\lesssim &
\norm{ \norm{\chi u}_{H^{\nu+\frac{1}{2}}}
\norm{\chi u}_{H^\frac{3}{2}}^2 }_{L^1_I}\\
\lesssim & \left\{\begin{aligned}
&\int_I{
\frac{e^{\left(1+(\kappa-\frac{1}{2})\right)\frac{m}{b(\tau)}}}
{\lambda^{3-2(\kappa-\frac{1}{2})}(\tau)}
e^{2\frac{2m+\pi}{b(\tau)}}\,d\,\tau}
&&\text{ for } \kappa < 2,\\
&\int_I{
e^{3\frac{2m+\pi}{b(\tau)}}\,d\,\tau}
&&\text{ for } \kappa = 2.
\end{aligned}\right.
\end{aligned}\end{equation}
Apply Lemma \ref{Lemma-lambdaIntegralCrude} for $\kappa \geq \frac{3}{2}$, Corollary \ref{Corollary-lambdaIntegral2version2} for $\kappa=1$, and Lemma \ref{Lemma-lambdaIntegral} for $\kappa = \frac{1}{2}$. Note that the result of equation (\ref{Proof-OriginHLower-useMoser}) is an entire order better in $\frac{1}{\lambda}$ than necessary.
\end{proof}
This completes the proof of Proposition \ref{Prop-Improv}.
\section{Proof of Theorem \ref{Thm-MainResult}}
\label{Section-FinalProof}
\begin{proof}[Proof of norm growth (\ref{Thm-MainResult-LogLog}), (\ref{Thm-MainResult-LogLogHigher})]
From Proposition \ref{Prop-Improv} we have that $T_{hyp} = T_{max}$, and from (\ref{Eqn-t1Small}) we have blowup in finite time. By the failure of local wellposedness we have that $\lambda(t) \to 0$ as $t \rightarrow T_{max}$. Recall the approximate dynamics of $\lambda$, equation (\ref{Eqn-lambda-prelimDynamics}), which with the control on $b$ implies in particular that $\abs{\frac{\lambda_s}{\lambda}} < 1$ on $[s_0,s_{max})$, which easily integrates to,
\begin{equation}\label{Eqn-sIsInfty}\begin{aligned}
\abs{\log \lambda(s)} \lesssim 1+s
&& \Longrightarrow
&& s_{max} = +\infty.
\end{aligned}\end{equation}
By direct calculation and a change of variable,
\[\begin{aligned}
-\partial_t\left(\lambda^2\log\abs{\log\lambda}\right) =
&-\frac{\lambda_s}{\lambda}\log\abs{\log\lambda}
\left(2 + \frac{1}{\abs{\log\lambda}\log\abs{\log\lambda}}\right).
\end{aligned}\]
From the approximate dynamics (\ref{Eqn-lambda-prelimDynamics}),
$
\frac{b}{2} \leq -\frac{\lambda_s}{\lambda} \leq 2b,
$
and so with the log-log rate {\bf H1.3} we have proven that, for some universal constant $C>0$ and all $t \in [0,T_{max})$,
\begin{equation}\label{Proof-MainResult-eqn1}
\frac{1}{C} \leq -\partial_t\left(\lambda^2\log\abs{\log\lambda}\right) \leq C.
\end{equation}
For all $t\in[0,T_{max})$, integrate equation (\ref{Proof-MainResult-eqn1}). Since $\lambda$ is very small we may estimate,
\begin{equation}\label{Proof-MainResult-BoundForLambda}\begin{aligned}
\frac{1}{C}\left(\frac{T_{max}-t}{\log\abs{\log(T_{max}-t)}}\right)^\frac{1}{2}
\leq \lambda(t)
\leq C \left(\frac{T_{max}-t}{\log\abs{\log(T_{max}-t)}}\right)^\frac{1}{2}.
\end{aligned}\end{equation}
We do not prove the exact value of the constant in equation (\ref{Thm-MainResult-LogLog}) - see \cite[Proposition 6]{MR-SharpLowerL2Critical-06}.
Finally, we conclude that equation (\ref{Thm-MainResult-LogLogHigher}) follows from the log-log relationship {\bf H1.3}, higher-order norm control {\bf H2.1}, and from $m > 0$ small.
As an aside, recall that $\frac{ds}{dt} = \frac{1}{\lambda^2}$, so that with (\ref{Proof-MainResult-BoundForLambda}) one would conclude,
\begin{equation}\label{Proof-MainResult-BoundForS}
\frac{1}{C}\abs{\log(T_{max}-t)} \leq s(t) \leq C\abs{\log(T_{max}-t)}.
\end{equation}
Then from the explicit lower and upper bounds for $b$, equations (\ref{Eqn-bLowerBound}) and (\ref{Eqn-bUpperBound}),
\begin{equation}\label{Proof-MainResult-BoundForB}
\frac{1}{C \log\abs{\log(T_{max}-t)}}
\leq b(t)
\leq \frac{C}{\log\abs{\log(T_{max}-t)}}.
\end{equation}
\end{proof}
\begin{proof}[Proof of stable locus of concentration, (\ref{Thm-MainResult-RZ})]
The preliminary estimate (\ref{Eqn-prelimGamma+REst}) implies in particular that $\abs{\frac{\partial_s(r,z)}{\lambda}} < 1$ on $[s_0,s_1)$. Then by change of variable, equation (\ref{Proof-MainResult-BoundForLambda}) and the bound on $T_{max}$, equation (\ref{Eqn-t1Small}),
\begin{equation}\label{Proof-MainResult-RZ}
\int_0^{T_{max}}{\abs{\partial_t(r,z)}\,dt} < \int_0^{T_{max}}{\frac{1}{\lambda(t)}\,dt} < \delta(\alpha^*).
\end{equation}
Equation (\ref{Thm-MainResult-RZ}) follows from choice of initial data {\bf C1.1}
\end{proof}
\begin{proof}[Proof of regularity away from singular ring, (\ref{Thm-MainResult-SingOnRing})]
Given $R>0$, define $\chi_R$ to be a suitable modification of $\chi$ (\ref{DefnEqn-Chi}), equal to one for $\abs{(r,z)-(r_{max},z_{max})} > R$.
Choose some $t(R)\in[0,T_{max})$ such that,
\begin{equation}\label{Proof-MainResult-BootstrapInitial}\begin{aligned}
A(t)\lambda(t) + \abs{(r(t),z(t))-(r_{max},z_{max})} \ll R
&& \text{ for all } t\in[t(R),T_{max}),
\end{aligned} \end{equation}
and hence $\chi_R u = \chi_R\widetilde{u}$ for all $t\in[t(R),T_{max})$.
Hypothesize $t_3\in(t(R),T_{max}]$ to be the largest value such that,
\begin{equation}\label{Proof-MainResult-Hcrit}\begin{aligned}
\norm{\chi_R u(t)}_{H^1} < 2\norm{\chi_R u(t(R))}_{H^1}
&& \text{ for all } t\in[t(R),t_3).
\end{aligned}\end{equation}
This choice of $t_3>t(R)$ is possible since $u(t)$ is strongly continuous in $H^1$ at time $t(R) < T_{max}$.
With interpolation, (\ref{Proof-MainResult-Hcrit})
replaces bootstrap hypotheses {\bf H2.2} and {\bf H2.3}. Repeating the arguments of Chapter \ref{Section-BootAtInfty} proves $t_3 = T_{max}$ and,
\begin{equation}\label{Proof-MainResult-SingOnRing}\begin{aligned}
\norm{\widetilde{u}(t)}_{H^{1}\left(\abs{(r,z)-(r_{max},z_{max})}>R\right)} < C(R)
&& \text{ for all } t \in[0,T_{max}).
\end{aligned}\end{equation}
\end{proof}
\begin{proof}[Proof of mass concentration, (\ref{Thm-MainResult-L2})]
Let $R> 0$. To begin we will prove there exists a residual profile in $L^2$ away from the singular ring,
\begin{equation}\label{Proof-MainResult-eqn3}\begin{aligned}
\tilde{u}(t) \to u^*
&& \text{ in } && L^2_x\left(\abs{(r,z)-(r_{max},z_{max})} \geq R\right)
&& \text{ as } && t \rightarrow T_{max}.
\end{aligned}\end{equation}
Then to establish equation (\ref{Thm-MainResult-L2}) we will prove $u* \in L^2(\real^{3})$,
\begin{equation}\label{Proof-MainResult-eqn3.2}\begin{aligned}
u^*\in L^2(\real^{3})
&& \text{ and }
&&\int{\abs{u^*}^2} = \lim_{t\rightarrow T_{max}}\int{\abs{\tilde{u}(t)}^2}.
\end{aligned}\end{equation}
Let $\epsilon_0 > 0$ be arbitrary. Due to equation (\ref{Eqn-epsIntegral}), we may choose $t(R)<T_{max}$ such that both,
\begin{equation}\label{Proof-MainResult-eqn4}\begin{aligned}
T_{max}-t(R) < \frac{\epsilon_0}{1+C\left(\frac{R}{4}\right)}
&&\text{ and }
&&\int_{t(R)}^{T_{max}}{\int{\abs{\grad\widetilde{u}}^2\,dx}\,dt} < \epsilon_0,
\end{aligned}\end{equation}
where $C(\frac{R}{4})$ is the constant from equation (\ref{Proof-MainResult-SingOnRing}). We may assume that, for $t\in[t(R),T_{max})$, $u(t) = \tilde{u}$ on $\left\{\abs{(r,z)-(r_{max},z_{max})} > \frac{R}{4}\right\}$.
For parameter $\tau > 0$, to be fixed later, we denote,
\begin{equation}\label{Proof-MainResult-DefnEqn-vTau}
v^\tau(t,x) = u(t+\tau,x) - u(t,x).
\end{equation}
Since $t(R) < T_{max}$, $u(t)$ is strongly continuous in $L^2$ at time $t(R)$. Thus, there exists $\tau_0$ such that,
\begin{equation}\label{Proof-MainResult-eqn5}\begin{aligned}
\int{\abs{v^\tau(t(R))}^2\,dx} < \epsilon_0
&& \text{ for all } \tau \in [0,\tau_0].
\end{aligned}\end{equation}
Denote $\phi_R$ a smooth cutoff function analogous to $\phi_\infty$ of equation (\ref{DefnEqn-phiInfty}),
\begin{equation}\label{Proof-MainResult-DefnEqn-phiR}
\phi_R(r,z) = \phi_\infty^4\left(\frac{(r,z) - (r_{max},z_{max})}{R}\right).
\end{equation}
By direct calculation,
\begin{equation}\label{Proof-MainResult-eqn7}\begin{aligned}
\frac{1}{2}\partial_t\left(\int{\phi_R\abs{v^\tau}^2}\right)
=& Im\left(\int{\grad\phi_R\cdot\grad v^\tau \overline{v^\tau}\,dx}\right)\\
&+ Im\left(\int{\phi_R v^\tau\left(\overline{u\abs{u}^2(t+\tau) - u\abs{u}^2(t)}\right)\,dx}\right).
\end{aligned}\end{equation}
Regarding the first RH term of (\ref{Proof-MainResult-eqn7}), from H\"older and our choice of $t(R)$ we have that,
\begin{equation}\label{Proof-MainResult-eqn8}
\int_{t(R)}^{T_{max}}\abs{
Im\left(\int{\grad\phi_R\cdot\grad v^\tau \overline{v^\tau}\,dx}\right)
\,dt}
\leq C \left(\int_{t(R)}^{T_{max}}{1^2\,dt}\right)^\frac{1}{2}\epsilon_0^\frac{1}{2}
< C\epsilon_0.
\end{equation}
Regarding the second RHS term of (\ref{Proof-MainResult-eqn7}), by homogeneity,
\begin{equation}\label{Proof-MainResult-eqn9}
\abs{\phi_R v^\tau\left(\overline{u\abs{u}^2(t+\tau) - u\abs{u}^2(t)}\right)}
\leq C\left(\abs{\phi_R^\frac{1}{4}u(t+\tau)}^4 + \abs{\phi_R^\frac{1}{4}u(t)}^4\right).
\end{equation}
Then, as we did in proving estimate (\ref{Eqn-4thOrderEst}), apply the Sobolev embedding $H^{\frac{3}{4}} \hookrightarrow L^4(\real^{3})$ and interpolate,
$\int{\abs{\phi_R^\frac{1}{4}u}^4} \leq C\norm{\phi_R^\frac{1}{4}u}_{H^\frac{1}{2}}^2
\norm{\phi_R^\frac{1}{4}u}_{H^1}^2$.
By the uniform control of $H^\frac{1}{2}$, equation (\ref{Proof-MainResult-SingOnRing}), and our choice of $t(R)$,
\begin{equation}\label{Proof-MainResult-eqn10}
\int_{t(R)}^{T_{max}}{\abs{\phi_R v^\tau\left(\overline{u\abs{u}^2(t+\tau) - u\abs{u}^2(t)}\right)}\,dt}
\leq C\epsilon_0.
\end{equation}
Through the integration of equation (\ref{Proof-MainResult-eqn7}) we have proven,
\begin{equation}\label{Proof-MainResult-eqn6}\begin{aligned}
\int{\phi_R\abs{v^\tau(t)}^2\,dx} < C \epsilon_0
&& \text{ for all } \tau \in [0,\tau_0]
\text{ and } t\in[t(R),T_{max}-\tau).
\end{aligned}\end{equation}
This shows that $\widetilde{u}$ is Cauchy, which proves (\ref{Proof-MainResult-eqn3}). We now turn our attention to (\ref{Proof-MainResult-eqn3.2}). Denote the thickness of the toroidal support of the singular profile and radiation by,
\begin{equation}\label{Proof-MainResult-DefnEqn-Rt}
R(t) = A(t)\lambda(t).
\end{equation}
Recall the definition of $A(t)$, equation (\ref{DefnEqn-A}). By the log-log rate {\bf H1.3}, $\lambda(t) \to 0$ implies that $A(t) \to 0$ and in particular,
\begin{equation}\label{Proof-MainResult-BoundForA}
A(t) \leq \frac{1}{\abs{\log(T_{max}-t)}^C}.
\end{equation}
Consider now $\phi_{R(t),\tau} = \phi_{\infty}^4\left(\frac{(r,z)-(r(\tau),z(\tau))}{R(t)}\right)$, a family of time-variable cutoffs similar to $\phi_{R(t)}$. For fixed time $t<T_{max}$ we calculate directly that,
\begin{equation}\label{Proof-MainResult-eqn11}
\begin{aligned}
\frac{1}{2}\partial_\tau\left(\int{\phi_{R(t),\tau}\abs{u(\tau)}^2\,dx}\right)
=&
\frac{1}{R(t)}Im\left(\int{\grad_x\phi_{R(t),\tau}\cdot\grad_xu(\tau)\overline{u(\tau)}\,dx}\right)\\
&-\frac{1}{2R(t)}\int{\partial_\tau(r(\tau),z(\tau))
\cdot\grad_x\phi_{R(t),\tau}\abs{u(\tau)}^2\,dx},
\end{aligned}\end{equation}
where we use $\grad_x\phi_{R(t),\tau}$ to denote $\left.\grad_y\phi_\infty^4(y)\right|_{y=\frac{(r,z)-(r(\tau),z(\tau))}{R(t)}}$.
Regarding the first RH line of (\ref{Proof-MainResult-eqn11}),
\[
\abs{\frac{1}{R(t)}
Im\left(\int{\grad_x\phi_{R(t),\tau}\cdot\grad_xu(\tau)\overline{u(\tau)}\,dx}\right)}
\lesssim \frac{1}{R(t)}\norm{u(\tau)}_{H^1} \lesssim \frac{1}{A(t)\lambda(t)\lambda(\tau)}.
\]
Regarding the second RH line of (\ref{Proof-MainResult-eqn11}),
apply the preliminary estimate (\ref{Eqn-prelimGamma+REst}),
\[
\abs{\frac{1}{2R(t)}\int{\partial_\tau(r(\tau),z(\tau))
\cdot\grad_x\phi_{R(t),\tau}\abs{u(\tau)}^2\,dx}}
\lesssim \frac{1}{A(t)\lambda(t)\lambda(\tau)}\int{\abs{u_0}^2}.
\]
Integrate (\ref{Proof-MainResult-eqn11}) in $\tau$, and apply the bounds for $A$ and $\lambda$ from equations (\ref{Proof-MainResult-BoundForA}) and (\ref{Proof-MainResult-BoundForLambda}),
\begin{multline}\label{Proof-MainResult-eqn12}
\abs{\int{\phi_{R(t),T_{max}}\abs{u^*}^2\,dx} - \int{\phi_{R(t),t}\abs{u(t)}^2\,dx}}\\
\begin{aligned}
&\leq C \frac{1}{A(t)\lambda(t)}\int_{t}^{T_{max}}{\frac{1}{\lambda(\tau)}\,d\tau}\\
&\leq \frac{C}{\abs{\log(T_{max}-t)}^C}\left(\frac{\log\abs{\log(T_{max}-t)}}{T_{max}-t}\right)^\frac{1}{2}
\int_t^{T_{max}}{\left(\frac{\log\abs{\log(T_{max}-\tau)}}{T_{max}-\tau}\right)^\frac{1}{2}\,d\tau}\\
&\leq \frac{1}{\abs{\log(T_{max}-t)}^\frac{C}{2}}.
\end{aligned}
\end{multline}
The final inequality relied upon $T_{max}-t < T_{max} < \alpha^*$, equation (\ref{Eqn-t1Small}), both to approximate the integral and then to approximate $C\log\abs{\log(T_{max}-t)} < \abs{\log(T_{max}-t)}^\frac{C}{2}$.
\begin{comment}
Rigourously, to perform the integral substitute $u = \log\abs{\log(T_{max}-t)}$ to get,
\[
\int_{\log\abs{\log(T_{max}-t)}}^{\infty}{\sqrt{ue^{e^u}}\frac{d}{du}\left(e^{-e^u}\right)\,du},
\]
Then by parts we have,
\[
2\sqrt{\log\abs{\log(T_{max}-t)}\,(T_{max}-t)} + \int_{\log\abs{\log(T_{max}-t)}}^{\infty}{\frac{e^{-\frac{1}{2}e^u}}{\sqrt{u}}\,du}.
\]
This final integral is clearly much less than what we started with. So we may approximate with $(2+\delta)\sqrt{\log\abs{\log(T_{max}-t)}\,(T_{max}-t)}$.
\end{comment}
Taking the limit $t \rightarrow T_{max}$ we see that,
\begin{equation}\label{Proof-MainResult-eqn13}
\int{\abs{u^*}^2} = \lim_{t\to T_{max}}\int{\phi_{R(t),t}\abs{u(t)}^2}.
\end{equation}
From the definition of $R(t)$ and the geometric decomposition,
\[
\lim_{t\to T_{max}}\int{\phi_{R(t),t}\abs{u(t)}^2}
=\lim_{t\to T_{max}}\int{\phi_{R(t),t}\abs{\widetilde{u}(t)}^2}
=\lim_{t\to T_{max}}\int{\abs{\widetilde{u}(t)}^2},
\]
which proves that the limit in (\ref{Proof-MainResult-eqn13}) exists and establishes equation (\ref{Proof-MainResult-eqn3.2}). This completes the proof of equation (\ref{Thm-MainResult-L2}).
\end{proof}
\begin{remark}[Consistency with $u^*\notin H^1$]
\label{Remark-ProfileNonH1}
By repeating the proof of {\bf I2.3}, we expect that following the proof of (\ref{Proof-MainResult-eqn3}) it could be shown that $\tilde{u}(t) \to u^*$ in
$H^1\left(\abs{(r,z)-(r_{max},z_{max})}\geq R\right)$.
Nevertheless, an attempt to prove a version of (\ref{Proof-MainResult-eqn3.2}) in $H^1$ will fail. Indeed, the second RH line of equation (\ref{Proof-MainResult-eqn11}) would require a bound for $\abs{\grad u(\tau)}$ on the support of $\grad\phi$, with
nothing to take the role mass conservation.
\end{remark}
|
\section{Introduction}
It is well known that F. Zwicky introduced the concept of dark
matter to account for the anomalous rotation curves of the galaxies
\cite{narlikarcos,tduniv}. The problem was that according to the
usual Newtonian Dynamics the velocities of the stars at the edges of
galaxies should fall with distance as in Keplarian orbits, roughly
according to
\begin{equation}
v \approx \sqrt{\frac{GM}{r}}\label{e1}
\end{equation}
where $M$ is the mass of the galaxy, $r$ the distance from the
centre of the galaxy of the outlying star and $v$ the tangential
velocity of the star. Observations however indicated that the
velocity curves flatten out, rather than follow the law (\ref{e1}).
This necessitated the introduction of the concept of dark matter
which would take care of the discrepancy without modifying Newtonian
dynamics. However even after nearly eight decades, dark matter has
not been detected, even though there have been any number of
candidates proposed for this, for example SUSY particles, massive
neutrinos, undetectable brown dwarf stars,
even black holes and so on.\\
Very recent developments are even more startling. These concern the
rotating dwarf galaxies, which are satellites of the Milky Way
\cite{metz1,metz2}. These studies throw up a big puzzle. On the one
hand these dwarf satellites cannot contain any dark matter and on
the other hand the stars in the satellite galaxies are observed to
be moving much faster than predicted by Newtonian dynamics, exactly
as in the case of the galaxies themselves. Metz, Kroupa, Theis,
Hensler and Jerjen conclude that the only explanation lies in
rejecting dark matter and Newtonian gravitation. Indeed a well known
Astrophysicist, R. Sanders from the University of Groningen
commenting on these studies notes \cite{physorg}, ``The authors of
this paper make a strong argument. Their result is entirely
consistent with the expectations of modified Newtonian dynamics
(MOND), but completely opposite to the predictions of the dark
matter hypothesis. Rarely is
an observational test so definite." Moreover Vahe Petrosian of the Kavli Institute in the February 10 issue of the Astrophysical Journal, rules out dark matter on the basis of studying the interaction of electrons at the galactic edge with starlight.\\
In this note we point out that this could indeed be so, though not
via Milgrom's ad hoc modified dynamics \cite{mil1,mil2,tduniv},
according to which a test particle at a distance $r$ from a large
mass $M$ is subject to the acceleration $a$ given by
\begin{equation}
a^2/a_0 = MGr^{-2},\label{3em1}
\end{equation}
where $a_0$ is an acceleration such that standard Newtonian dynamics
is a good approximation only for accelerations much larger than
$a_0$. The above equation however would be true when $a$ is much
less than $a_0$. Both the statements in (\ref{3em1}) can be combined
in the heuristic relation
\begin{equation}
\mu (a/a_0) a = MGr^{-2}\label{3em2}
\end{equation}
In (\ref{3em2}) $\mu(x) \approx 1$ when $x >> 1, \, \mbox{and}\,
\mu(x) \approx x$ when $x << 1$. It must be stressed that
(\ref{3em1}) or (\ref{3em2}) are not deduced from any theory, but
rather are an ad hoc prescription to explain observations.
Interestingly it must be mentioned that most of the implications of
Modified Newtonian Dynamics or MOND do not
depend strongly on the exact form of $\mu$.\\
It can then be shown that the problem of galactic velocities is now
solved \cite{mil1,mil2,mil3,mil4,mil5}. Nevertheless, most
physicists are not comfortable with MOND because of the ad hoc
nature of (\ref{3em1}) and (\ref{3em2}).
\section{Varying $G$ Dynamics}
We now come to the cosmological model described by the author in
1997 (Cf.ref.\cite{ijmpa1998,tduniv} and several references
therein), in which the universe, under the influence of dark energy
would be accelerating with a small acceleration. Several other
astrophysical relations, some of them hitherto inexplicable such as
the Weinberg formula giving the pion mass in terms of the Hubble
constant were also deduced in this model (Cf.also ref.\cite{cu} and
references therein). While all this was exactly opposite to the then
established theory, it is well known that the picture was
observationally confirmed soon thereafter through the work of
Perlmutter and others (Cf.ref.\cite{cu}). Interestingly, in this
model Newton's
gravitational constant varied inversely with time.\\
Cosmologies with time varying $G$ have been considered in the past,
for example in the Brans-Dicke theory or in the Dirac large number
theory or by Hoyle \cite{barrowparsons,narfpl,narburbridge,5,6}. In
the case of the Dirac cosmology, the motivation was Dirac's
observation that the supposedly large number coincidences involving
$N \sim 10^{80}$, the number of elementary particles in the universe
had an underlying message if it is recognized that
\begin{equation}
\sqrt{N} \propto T\label{3ea1}
\end{equation}
where $T$ is the age of the universe. Equation (\ref{3ea1}) too
leads to a $G$ decreasing inversely
with time as we will now show. We follow a route slightly different from that of Dirac.\\
From (\ref{3ea1}) it can easily be seen that
\begin{equation}
T = \sqrt{N} \tau\label{5}
\end{equation}
where $\tau$ is a typical Compton time of an elementary particle
$\sim 10^{-23}secs$, because $T$, the present age of the universe is
$\sim 10^{17}secs$. We also use the following relation for a
uniformly expanding Friedman Universe
\begin{equation}
\dot{R}^2 = \frac{8 \pi}{3} G \, R^2 \rho\label{6}
\end{equation}
where $R$ is the radius of the universe and $\rho$ its density. We
remember that
\begin{equation}
\rho = \frac{3M}{4 \pi R^3} \, \mbox{and} \, M = Nm\label{7}
\end{equation}
where $M$ is the mass of the universe, and $m$ is the mass of an
elementary particle $\sim 10^{-25}gm$
(Cf.ref.\cite{weinberggravcos}).\\
Use of (\ref{7}) in (\ref{6}) leads to another well known relation
\cite{ruffini}
\begin{equation}
R = \frac{GM}{c^2}\label{8}
\end{equation}
because $\dot{R} = c$. Further dividing both sides of (\ref{5}) by
$c$ we get the famous Weyl-Eddington relation
\begin{equation}
R = \sqrt{N} l\label{9}
\end{equation}
where $l = \tau /c$ is a typical Compton length $\sim 10^{-13}cms$.\\
Use of (\ref{7}) and (\ref{9}) in (\ref{8}) now leads to
\begin{equation}
G = \frac{c^2 l}{\sqrt{N}m} = \left(\frac{c^2 l \tau}{m}\right)
\cdot \frac{1}{T} \equiv \frac{G_0}{T}\label{10}
\end{equation}
Equation (\ref{10}) gives the above stated inverse dependence of the
gravitational constant $G$ on time, which Dirac obtained. On the
other hand this same relation was obtained by a different route in
the author's dark energy -- fluctuations cosmology in 1997. This
work, particularly in the context of the Planck scale has been there
for many years in the literature (Cf.\cite{cu,tduniv,uof} and
references therein). Suffice to say that all the supposedly so
called accidental Large Number Relations like (\ref{9}) as also the
inexplicable Weinberg formula which relates the Hubble constant to
the mass of a pion, follow as deductions in this
cosmology. The above references give a comprehensive picture.\\
The Brans-Dicke cosmology arose from the work of Jordan who was
motivated by Dirac's ideas to try and modify General Relativity
suitably. In this scheme the variation of $G$ could be obtained from
a scalar field $\phi$ which would satisfy a conservation law. This
scalar tensor gravity theory was further developed by Brans and
Dicke, in which $G$ was inversely proportional to the variable field
$\phi$. (It may be mentioned
that more recently the ideas of Brans and Dicke have been further generalized.)\\
In the Hoyle-Narlikar steady state model, it was assumed that in the
Machian sense the inertia of a particle originates from the rest of
the matter present in the universe. This again leads to a variable
$G$. The above references give further details of these various
schemes and their shortcomings which have
lead to their falling out of favour.\\
In any case, our starting point is, equation (\ref{10}) where $T$ is
time (the age of the universe) and $G_0$ is a constant. Furthermore,
other routine effects like the precession of the perihelion of
Mercury and the bending of light and so on are also explained with
(\ref{10}) as will be briefly discussed below. We will also see
that there is observational evidence for (\ref{10}) (Cf. also
\cite{uzan} which described various observational evidences for the
variation of $G$, for example from solar system observations, from
cosmological observations and even from the
palaeontological studies point of view).\\
With this background, we now mention some further tests for equation
(\ref{10}).\\
This could explain the other General Relativistic effects like the
shortening of the period of binary pulsars and so on
(Cf.ref.\cite{cu,tduniv,bgsnc115b,bgsfpl} and other references
therein). Moreover, we could now also explain, the otherwise
inexplicable anomalous acceleration of the Pioneer space crafts
(Cf.ref.\cite{tduniv} for details). We will briefly revisit some of these effects later.\\
We now come to the problem of galactic rotational curves mentioned
earlier (cf.ref.\cite{narlikarcos}). We would expect, on the basis
of straightforward dynamics that the rotational velocities at the
edges of galaxies would fall off according to
\begin{equation}
v^2 \approx \frac{GM}{r}\label{3ey33}
\end{equation}
which is (\ref{e1}). However it is found that the velocities tend to
a constant value,
\begin{equation}
v \sim 300km/sec\label{3ey34}
\end{equation}
This, as noted, has lead to the postulation of the as yet undetected
additional matter alluded to, the so called dark matter.(However for
an alternative view point Cf.\cite{sivaramfpl93}). We observe that
(\ref{10}) can be written for an increase $t$,in time, small
compared to the age of the universe, now written as $t_0$
\begin{equation}
G = \frac{G_0}{t_0 + t} = \frac{G_0}{t_0} \left(1 -
\frac{t}{t_0}\right)\label{3.17}
\end{equation}
Using (\ref{3.17}), let us consider the gravitational potential
energy $V$ between two masses, $m_1$ and $m_2$ by:
\begin{equation}
V = \frac{G m_1 m_2}{r_0} = \frac{G_0}{t_0} \cdot
\frac{m_1m_2}{r_0}\label{14}
\end{equation}
After a time $t$ this would be, by (\ref{3.17}),
\begin{equation}
V = \frac{G_0}{t_0} \left(1 - \frac{t}{t_0}\right)
\frac{m_1m_2}{r}\label{15}
\end{equation}
Equating (\ref{14}) and (\ref{15}) we get,
\begin{equation}
r = r_0 \left(\frac{t_0}{t_0 + t}\right)\label{3.18}
\end{equation}
The relation (\ref{3.18}) has been deduced by a different route by
Narlikar \cite{narlikarcos}. From (\ref{3.18}) it easily follows
that,
\begin{equation}
a \equiv (\ddot{r}_{o} - \ddot{r}) \approx \frac{1}{t_o}
(t\ddot{r_o} + 2\dot r_o) \approx -2 \frac{r_o}{t^2_o}\label{3ey35}
\end{equation}
as we are considering intervals $t$ small compared to the age of the
universe and nearly circular orbits. In (\ref{3ey35}), $a$ or the
left side of (\ref{3ey35}) gives the new extra effect due to
(\ref{3.17}) and (\ref{3.18}), this being a departure from the usual
Newtonian gravitation. Equation (\ref{3ey35}) shows
(Cf.ref\cite{bgsnc115b} also) that there is an anomalous inward
acceleration, as if there is
an extra attractive force, or an additional central mass.\\
So, introducing the extra acceleration (\ref{3ey35}), we get,
\begin{equation}
\frac{GMm}{r^2} + \frac{2mr}{t^2_o} \approx
\frac{mv^2}{r}\label{3ey36}
\end{equation}
From (\ref{3ey36}) it follows that
\begin{equation}
v \approx \left(\frac{2r^2}{t^2_o} + \frac{GM}{r}\right)^{1/2}
\label{3ey37}
\end{equation}
So (\ref{3ey37}) replaces (\ref{e1}) in this model. This shows that
as long as
\begin{equation}
\frac{2r^2}{t_0^2} < < \frac{GM}{r},\label{20}
\end{equation}
Newtonian dynamics holds. But when the first term on the left side
of (\ref{20}) becomes of the order of the second (or greater), the
new dynamical effects come in.\\
For example from (\ref{3ey37}) it is easily seen that at distances
well within the edge of a typical galaxy, that is $r < 10^{23}cms$
the usual equation (\ref{3ey33}) holds but as we reach the edge and
beyond, that is for $r \geq 10^{24}cms$ we have $v \sim 10^7 cms$
per second, in agreement with (\ref{3ey34}). In fact as can be seen
from (\ref{3ey37}), the first term in the square root has an extra
contribution (due to the varying $G$) which exceeds the second term
as we approach the galactic edge, as if there is an extra mass, that
much more.\\
We would like to stress that the same conclusions will apply to the
latest observations of the satellite galaxies (without requiring any
dark matter). Let us for example consider the Megallanic clouds
\cite{stave}. In this case, as we approach their edges, the first
term within the square root on the right side of (\ref{3ey37}) or
the left term of (\ref{20}) already becomes of the order of the
second term, leading to the new non Newtonian effects.
\section{Remarks}
We have already noted that the varying $G$ dynamics given in
(\ref{10}) explains all the General Relativistic effects. This work
has been available in the literature for many years
(Cf.\cite{bgsnc115b,cu,uof,tduniv} for full details and further
references). However we repeat some examples briefly to give an idea.\\
We could explain the correct gravitational bending of light. Infact
in Newtonian theory too we obtain the bending of light, though the
amount is half that predicted by General
Relativity\cite{narlikarcos,denman,silverman,brill}. In the
Newtonian theory we can obtain the bending from the well known
orbital equations (Cf.also\cite{gold}),
\begin{equation}
\frac{1}{r} = \frac{GM}{L^2} (1+ecos\Theta)\label{3ey25}
\end{equation}
where $M$ is the mass of the central object, $L$ is the angular
momentum per unit mass, which in our case is $bc$, $b$ being the
impact parameter or minimum approach distance of light to the
object, and $e$ the eccentricity of the trajectory is given by
\begin{equation}
e^2 = 1+ \frac{c^2L^2}{G^2M^2}\label{3ey26}
\end{equation}
For the deflection of light $\alpha$, if we substitute $r = \pm
\infty$, and then use (\ref{3ey26}) we get
\begin{equation}
\alpha = \frac{2GM}{bc^2}\label{3ey27}
\end{equation}
This is half the General
Relativistic value.\\
We now note that the effect of time variation of $r$ is given by
equation (\ref{3.18})(cf.ref.\cite{bgsnc115b}). Using this the well
known equation for the trajectory is given by,
\begin{equation}
u" + u = \frac{GM}{L^2} + u\frac{t}{t_0} + 0 \left (
\frac{t}{t_0}\right )^2\label{3ey28}
\end{equation}
where $u = \frac{1}{r}$ and primes denote differenciation with
respect to
$\Theta$.\\
The first term on the right hand side represents the Newtonian
contribution while the remaining terms are the contributions due to
(\ref{10}). The solution of (\ref{3ey28}) is given by
\begin{equation}
u = \frac{GM}{L^2} \left[ 1 + ecos\left\{
\left(1-\frac{t}{2t_0} \right) \Theta +
\omega\right\} \right]\label{3ey29}
\end{equation}
where $\omega$ is a constant of integration. Corresponding to
$-\infty < r < \infty$ in the Newtonian case we have in the present
case, $-t_0 < t < t_0$, where $t_0$ is large and infinite for
practical purposes. Accordingly the analogue of the reception of
light for the observer, viz., $r = + \infty$ in the Newtonian case
is obtained by taking $t = t_0$ in (\ref{3ey29}) which gives
\begin{equation}
u = \frac{GM}{L^2} + ecos \left(\frac{\Theta}{2} + \omega
\right)\label{3ey30}
\end{equation}
Let us compare (\ref{3ey30}) with the Newtonian solution. This shows
that the Newtonian $\Theta$ is replaced by $\frac{\Theta}{2}$,
whence the deflection obtained by equating the left side of
(\ref{3ey30}) to zero, is
\begin{equation}
cos \Theta \left(1-\frac{t}{2t_0}\right) = -\frac{1}{e}\label{3ey31}
\end{equation}
where $e$ is given by (\ref{3ey26}). The value of the deflection
from (\ref{3ey31}) is twice the Newtonian deflection given by
(\ref{3ey27}). That is the deflection $\alpha$ is now given not by
(\ref{3ey27}) but by the formula,
\begin{equation}
\alpha = \frac{4GM}{bc^2},\label{3ey32}
\end{equation}
The relation (\ref{3ey32}) is the correct observed value and is the
same as the General Relativistic formula which however is obtained
by a different
route \cite{brill,berg,lass}.\\
Next, we come to the inexplicable anomalous accelerations of the
Pioneer spacecrafts already alluded to, which have been observed by
J.D. Anderson and coworkers at the Jet Propulsion Laboratory for
well over a decade \cite{andersonphysrev,andersongrqc} and have
posed a puzzle. This can be explained in a simple way as follows: In
fact from the usual orbital equations we have \cite{bgstest}
$$v \dot {v} \approx -\frac{GM}{2t_0 r} (1 + e cos \Theta )-\frac{GM}{r^2} \dot {r}(1+e cos \Theta )$$
$v$ being the velocity of the spacecraft and $t$ is the time in
general. It must be observed that the first term on the right side
is the new effect due to (\ref{10}). There is now an anomalous
acceleration given by
$$a_r = \langle \dot v \rangle_{\mbox{anom}} = \frac{-GM}{2t r v} (1+e cos \Theta )$$
$$\approx -\frac{GM}{2t\lambda} (1+e)^3$$
where
$$\lambda = r^4 \dot \Theta^2$$
If we insert the values for the Pioneer spacecrafts we get
$$a_r \sim -10^{-7} cm/sec^2$$
This is the anomalous acceleration reported by Anderson and co-workers.\\
We will next deduce that equations like (\ref{3.18}) explains
correctly the observed decrease in the orbital period of the binary
pulsar $PSR\, 1913 + 16$, which has also been attributed to as yet
undetected gravitational waves \cite{davis}.\\
It may be observed that the energy $E$ of two masses $M$ and $m$ in
gravitational interaction at a distance $L$ is given by
\begin{equation}
E = \frac{GMm}{L} = \mbox{constant}\label{3ea3}
\end{equation}
We note that if this energy decreases by any mechanism, for example
by the emission of gravitational waves, or by the decrease of $G$,
then because of (\ref{3ea3}), there is a compensation by the
decrease in the orbital length and orbital period. This is the
standard General Relativistic explanation for the binary pulsar
$PSR\, 1913 + 16$. We will show that the same holds good, if we are
given instead, (\ref{10}). That is, we will not invoke gravitational
waves. In this case we have, from (\ref{3ea3})
\begin{equation}
\frac{\mu}{L} \equiv \frac{GMm}{L} = \mbox{const.}\label{3ea4}
\end{equation}
We can now write, for a time increase $t$,
\begin{equation}
\mu = \mu_0 - Kt\label{3ea5}
\end{equation}
where we have
\begin{equation}
K \equiv \dot{\mu}\label{3ea6}
\end{equation}
In (\ref{3ea6}) $\dot{\mu}$ can be taken to be a constant in view of
the fact that $G$ varies very slowly with time. Specifically we have
\begin{equation}
G (T + t) = G(T) -t \frac{G}{T} + \frac{t^2}{2} \frac{G}{T^2} +
\cdots \approx G(T) - t \frac{G}{T}\label{3ea7}
\end{equation}
where $T$ is the age of the universe and $t$ is an incremental time.
Whence using (\ref{3ea7}), $K$ in (\ref{3ea6}) is given by
$$K \propto \frac{G}{T}$$
and so
$$\dot{K} \sim \frac{G}{T^2} \approx 0$$
So (\ref{3ea4}) requires
$$L = L_0 (1 - \alpha K)$$
Whence on using (\ref{3ea5}) we get
\begin{equation}
\alpha = \frac{t}{\mu_0}\label{3ea8}
\end{equation}
Let us now consider $t$ to be the period of revolution in the case
of the binary pulsar. Using (\ref{3ea8}) it follows that
\begin{equation}
\delta L = - \frac{L_0tK}{\mu_0}\label{3ea9}
\end{equation}
We also know (Cf.ref.\cite{bgstest})
\begin{equation}
t = \frac{2\pi}{h} L^2 = \frac{2\pi}{\sqrt{\mu}}\label{3ea10}
\end{equation}
\begin{equation}
t^2 = \frac{4\pi^2 L^3}{\mu},\label{3ea11}
\end{equation}
$h$ being the usual unit angular momentum and $\mu$ has the units
$gm\, cm^4 sec^{-1}$. Using (\ref{3ea9}), (\ref{3ea10}) and
(\ref{3ea11}), a little manipulation gives
\begin{equation}
\delta t = - \frac{2t^2K}{\mu_0}\label{3ea12}
\end{equation}
(\ref{3ea9}) and (\ref{3ea12}) show that there is a decrease in the
size of the orbit, as also in the orbital period. Such a decrease in
the orbital period has been observed in the
case of binary pulsars in general \cite{davis,ohanian}.\\
Let us now apply the above considerations to the specific case of
the binary pulsar $PSR\, 1913 + 16$ observed by Taylor and coworkers
(Cf.ref.\cite{ohanian}). In this case it is known that, $t$ is 8
hours while $v$, the orbital speed is $3 \times 10^7 cms$ per
second. It is easy to calculate from the above
$$\mu_0 = 10^4 \times v^3 \sim 10^{26}$$
which gives $M \sim 10^{33}gms$, which of course agrees with
observation. Further we get using (\ref{3ea1}) and (\ref{3ea5})
\begin{equation}
\Delta t = \eta \times 10^{-5} sec/yr, \eta <\approx 8\label{3ea13}
\end{equation}
Indeed (\ref{3ea13}) is in good agreement with the carefully
observed value of
$\eta \approx 7.5$ (Cf.refs.\cite{davis,ohanian}).\\
It should also be remarked that in the case of gravitational
radiation, there are some objections relevant to the
calculation (Cf.ref.\cite{davis}).\\
Finally, we may point out that a similar shrinking in size with time
can be expected of galaxies themselves, and in general,
gravitationally bound systems.\\
To consider the above result in a more general context, we come back
to the well known orbital equation \cite{bgstest}
\begin{equation}
d^2 u/d\Theta^2 + u = \mu_0/h^2\label{3ea14}
\end{equation}
where $\mu_0 = GM$ and $u$ is the usual inverse of radial distance.\\
$M$ is the mass of the central object and $h = r^2 d\Theta / dt$ - a
constant. The solution of (\ref{3ea14}) is well known,
$$lu = 1 + ecos \Theta$$
where $l = h^2/\mu_0$.\\
It must be mentioned that in the above purely classical analysis,
there is no
precession of the perihelion.\\
\indent We now replace $\mu_0$ by $\mu$ and also assume $\mu$ to be
varying slowly because $G$ itself varies slowly and uniformly, as
noted earlier:
\begin{equation}
\dot{\mu} = d\mu / dt = K, \mbox{a \, constant}\label{3ea15}
\end{equation}
remembering that $\dot{K} \sim 0 (1/T^2)$ and so can be neglected.\\
\indent Using (\ref{3ea15}) in (\ref{3ea14}) and solving the orbital
equation (\ref{3ea14}), the solution can now be obtained as
\begin{equation}
u = 1/l + (e/l)cos\Theta + K l^2 \Theta/h^3 + Kl^2 e\Theta cos\Theta
/h^3\label{3ea16}
\end{equation}
Keeping terms up to the power of '$e$' and $(K/\mu_0 )^2$, the time
period '$\tau$' for one revolution is given to this order of
approximation by
\begin{equation}
\tau = 2\pi L^2/h\label{3ea17}
\end{equation}
From (\ref{3ea16})
\begin{equation}
L = l - \frac{Kl^4\Theta}{h^3}\label{3ea18}
\end{equation}
Substituting (\ref{3ea18}) in (\ref{3ea17}) we have
\begin{equation}
\tau = \frac{2\pi}{h} \left(l^2 -
\frac{2Kl^5\Theta}{h^3}\right)\label{3ea19}
\end{equation}
The second term in (\ref{3ea19}) represents the change in time
period for one revolution. The decrease of time period is given by
\begin{equation}
\delta \tau = 8\pi^2 l^3 K/\mu_0^2\label{3ea20}
\end{equation}
The second term in (\ref{3ea18}) indicates the decrease in latus-rectum.\\
\indent For one revolution the change of latus-rectum is given by
\begin{equation}
\delta l = 2\pi Kl^{2.5}/\mu^{1.5}_0\label{3ea21}
\end{equation}
In the solar system, we have,
$$K = 898800\, cm \, gm$$
Using $K$ and $\mu_0$ to find the change in time period and the
latus rectum in the varying $G$ case by substituting in
(\ref{3ea20}) and (\ref{3ea21}) respectively for Mercury we get
$$\delta T = 1.37 \times 10^{-5} sec/rev$$
\begin{equation}
\delta l = 4.54 cm/rev\label{3ea22}
\end{equation}
We observe that the equations (\ref{3ea20}), (\ref{3ea21}) or
(\ref{3ea22}) show a decrease in distance and in the time of
revolution. If we use for the planetary motion, the General
Relativistic analogue of (\ref{3ea14}), viz.,
$$\frac{d^2 u}{d\Theta^2} + u = \frac{\mu_0}{h^2} (1 + 3h^2 u^2),$$
then while we recover the precession of the perihelion of Mercury,
for example, there is no effect similar to (\ref{3ea20}),
(\ref{3ea21}) or (\ref{3ea22}). On the other hand this effect is
very minute-- just a few centimeters per year in the case of the
earth-- and only protracted careful observations can detect it.
Moreover these changes could also
be masked at least partly, by gravitational and other perturbations.\\
However as noted, the decrease of the period in (\ref{3ea20}) has
been observed in the case of Binary Pulsars.
\section{Conclusion}
The point is that the above varying $G$ scheme described in
(\ref{10}) or (\ref{3ey37}) reproduces all the effects of General
Relativity as noted above, as also the anomalous acceleration of the
Pioneer space crafts in addition to the conclusions of MOND
regarding an alternative for dark matter, and is applicable to the
latest observations of satellite galaxies. The satellite galaxy
rotation puzzle is thus resolved. \vspace{5 mm}
\begin{flushleft}
|
\section{Introduction}
\label{sec:intro}
Modeling social phenomena represents a major challenge that
has in recent years attracted a growing interest. Insight into the problem can be gained
by resorting, among others, to the so called {\em Agent Based Models}, an approach that is well
suited to bridge the gap between hypotheses
concerning the microscopic behavior of individual agents and the emergence
of collective phenomena in a population composed of many interacting
heterogeneous entities.
Constructing sound models deputed to return a reasonable approximation of the
scrutinized dynamics is a delicate operation, given the degree of arbitrariness
in assigning the rules that govern mutual interactions.
In the vast majority of cases, data are scarce and do not sufficiently constrain the model, hence the
provided answers can be questionable. Despite this intrinsic limitation,
it is however important to inspect the emerging dynamical properties of abstract models,
formulated so to incorporate the main distinctive traits of a social interaction scheme.
In this paper we aim at discussing one of such models, by combining analytical and numerical
techniques. In particular, we will focus on characterizing the evolution of the
underlying social network in terms of dynamical indicators.
It is nowadays well accepted that several social groups display
two main features: the {\em small world property}~\cite{WS1998} and the
presence of {\em weak ties}~\cite{Granovetter1983}. The first property implies that the network exhibits clear tendency to organize in large, densely connected, clusters.
As an example, the probability that two friends of mine are also, and independently, friends
to each other is large. Moreover, the shortest path between two generic individuals is
small as compared to the analogous distance computed for a
random network made of the same number of individuals and inter-links connections. This observation signals the
existence of short cuts in the social tissue. The second property
is related to the cohesion of the group which is mediated by small groups of well tied
elements, that are conversely weakly connected to other
groups. The skeleton of a social community is hence a hierarchy of subgroups.
A natural question arise on the ubiquity of the aforementioned peculariar aspects, distinctive traits of a real social networks:
how can they eventually emerge, starting from an finite group of randomly connected actors?
We here provide an answer to this question in the framework of a minimalistic opinion dynamics model,
which exploit an underlying substrate where opinions can flow. More specifically,
the network that defines the topological structure is imagined to evolve, coupled to the opinions and following a
specific set of rules: once two agents reach a compromise and share a common opinion, they also increase their
mutual degree of acquaintance, so strengthing the reciprocal link. In this respect, the model that we are shortly going to introduce
hypothesize a co-evolution of opinions and social structure, in the spirit of a genuine adaptive network~\cite{GrossBlasius,ZESM}.
Working within this framework, we will show that an initially generated random group, with
respect to both opinion and social ties, can evolve towards a final state where small worlds and weak ties effects are indeed
present. The results of this paper constitute the natural follow up of a series of
papers~\cite{pre,opi3,epjb}, where the time evolution of the
opinions and affinity, together with the fragmentation vs. polarization
phenomena, have been discussed.
Different continuous opinion dynamics models have been presented in
literature, see for instance~\cite{deffuant2000,galam2008}, dealing with the
general consensus problem. The aim is to shed light onto the assumptions
that can eventually yield to fixation, a final mono-clustered configuration where all
agents share the same belief, starting from an initial condition where the inspected population
is instead fragmented into several groups. In doing so, and in most cases, a
fixed network of interactions is a priori imposed~\cite{AD2004}, and the polarization dynamics studied under the constraint of
the imposed topology. At variance, and as previously remarked, we will instead allow the underlying network
to dynamically adjust in time, so modifying its initially imposed characteristics. Let us start by revisiting the
main ingredients of the model. A more detailed account can be found in ~\cite{pre}.
Consider a closed group of $N$ agents, each one
possessing its own opinion on a given subject. We
here represent the opinion of element $i$ as a continuous real variable
$O_i\in [0,1]$. Each agent is also characterized by its affinity score
with respect to the remaining $N-1$
agents, namely a vector $\alpha_{ij}$, whose entries are real number defined in the interval
$[0,1]$: the larger the value of the affinity $\alpha_{ij}$, the more reliable the relation of $i$ with the end node $j$.
Both opinion and affinity evolve in time because of binary encounters between
agents. It is likely that more interactions can potentially occur among
individuals that are more affine, as defined by the preceding indicator, or that
share a close opinion on a debated subject. Mathematically, these requirements can be accommodated for by
favoring the encounters between agents that minimizes the {\em social metric}
$D^t_{ij}=|\Delta O_{ij}^t|(1-\alpha_{ij}^t)+\mathcal{N}_j(0,\sigma)$, where
$\Delta O_{ij}^t=O_i^t-O_j^t$ is the opinions' difference of agents $i$ and $j$ at time $t$, and the last term is a stochastic
contribution, normally distributed with zero mean and
variance $\sigma$. For a more detailed analysis on
the interpretation of $\sigma$ as a {\em social temperature} responsible of a
increased mixing ability of the population, we refer
to~\cite{pre,opi3,epjb}.
Once two agents are selected for interaction they possibly update their opinions (if they are
affine enough) and/or change their affinities (if they have close enough
opinions), following:
\begin{equation}
\begin{cases}
O_i^{t+1} &= O_i^{t}- \frac{1}{2} \, \Delta O_{ij}^{t}\,
\Gamma_1\left(\alpha^t_{ij}\right) \\
\alpha_{ij}^{t+1} &= \alpha_{ij}^{t} + \alpha_{ij}^{t}
(1-\alpha_{ij}^{t}) \,\Gamma_2 \left(\Delta O^t_{ij}\right) \, ,
\end{cases}
\label{eq:themodel}
\end{equation}
being:
\begin{equation}
\Gamma_1 \left(x\right)= \frac{\tanh (\beta_1
(x-\alpha_c)) + 1 }{2} \quad\text{and}\quad
\Gamma_2 \left(x\right)= -\tanh(\beta_2 (|x| -
\Delta O_c)) \, ,
\end{equation}
two {\em activating functions} which formally reduce to step functions for large enough the values of
the parameters $\beta_1$ and $\beta_2$, as it is the case in the numerical simulations reported below.
Despite its simplicity the model exhibits an highly non linear dependence on
the involved parameters, $\alpha_c$, $\Delta O_c$ and $\sigma$, with a phase
transition between a polarized and fragmented dynamics~\cite{pre}.
A typical run for $N=100$ agents is reported in
the main panel of Fig.~\ref{fig:fig1}, for a choice of the parameters which
yields to a consensus state. The insets represent three successive time snapshots of
the underlying social network: The $N$ nodes are the individuals, while
the links are assigned based on the associated values of the affinity. The figures
respectively refer to a relatively early stage of the evolution $t=1000$, to an
intermediate time $t=5000$ and to the convergence time $T_c=10763$. Time is here calculated as the number
of iterations (not normalized with respect to $N$). The
corresponding three networks can be characterized using standard topological
indicators~\cite{AB2002,blmch2006} (see Table~\ref{tab:table1}), e.g. the mean
degree $<k>$, the network clustering coefficient $C$ and the average shortest path $<\ell >$. An
explicit definition of those quantities will be given below.
In the forthcoming discussion we will focus on the evolution of the network
topology, limited to a choice of the parameters that yield to a final mono cluster.
\begin{table}[ph]
\tbl{Topological indicators of the social networks presented in
Fig.~\ref{fig:fig1}. The mean degree $<k>$, the network clustering $C$ and
the average shortest path $<\ell >$ are
reported for the three time configurations depicted in the figure.}
{\begin{tabular}{@{}cccc@{}} \toprule
& $<k>(t)$ & $C(t)$ & $<\ell >(t)$ \\ \colrule
$t=1000$ & $0.073$ & $0.120$ & $3.292$ \\
$t=5000$ & $0.244$ & $0.337$ & $2.013$ \\
$t=T_c$ & $0.772$ & $0.594$ & $1.228$ \\ \botrule
\end{tabular}}
\label{tab:table1}
\end{table}
\begin{figure}[ph]
\centerline{\psfig{file=Fig1Net.eps,width=8cm}}
\vspace*{8pt}
\caption{Opinions as function of time. The run refers to
$\alpha_c=0.5$, $\Delta O_c=0.5$ and $\sigma =
0.01$. The underlying network is displayed at
different times, testifying on its natural tendency to evolve towards
a single cluster of affine individuals. Initial opinions are
uniformly distributed in the
interval $[0, 1]$, while $\alpha_{ij}^0$ are randomly assigned in $[0,1/2]$
with uniform distribution.}
\label{fig:fig1}
\end{figure}
\section{The social network}
\label{sec:evolsocnet}
The affinity matrix drives the interaction via the selection mechanism. It hence
can be interpreted as the {\em adjacency} matrix of the
underlying {\em social network}, i.e. the network of social ties that
influences
the exchange of opinions between acquaintances, as mediated by the encounters. Because the
affinity is a
dynamical variable of the model, we are actually focusing on
an {\em adaptive} social network~\cite{GrossBlasius,ZESM} : The network
topology influences in turn the dynamics of opinions, this latter providing a feedback on
the network itself and so modifying its topology. In other words, the evolution
of the topology is inherent
to the dynamics of the model because of the proposed self-consistent
formulation and not imposed a priori as an additional, external ingredient,
(as e.g. rewire and/or add/remove links according to a given
probability~\cite{HN,KB} once the state variables have been updated).
\begin{remark}[Weighted network]
Let us observe that the affinity assumes positive real values, hence we can
consider
a weighted social networks, where agents weigh the relationships. Alternatively, one can
introduce a cut-off parameter, $\alpha_f$: agents $i$ and $j$ are
socially linked if and only if the recorded relative affinity is large enough, meaning
$\alpha_{ij}>\alpha_f$. Roughly, the agent chooses its
closest friends among all his neighbors.
The first approach avoids the introduction of non--smooth functions and it is
suitable to carry on the analytical calculations. The latter results more straightforward
for numerical oriented applications.
\end{remark}
As anticipated, we are thus interested in analyzing the model, for a specific choice of the parameters, $\alpha_c$,
$\sigma$ and $\Delta O_c$, yielding to consensus, and studying the evolution
of the network topology, here analyzed via standard network indicators: the average value
of {\em weighted degree}, the {\em cluster coefficient} and the {\em averaged
shortest path}. These quantities will be quantified for (i)
a fixed population, monitoring their time dependence; (ii) as a function of the
population size, photographing the dynamics at convergence, namely when consensus has been reached.
\subsection{Time evolution of weighted degree}
\label{ssec:degree}
The simplest and the most intensively studied one--vertex (i.e. local)
characteristic is the node {\em degree}~\footnote{Let us observe that the
affinity may not be symmetric and thus the inspected social network will be directed.
One has thus to distinguish between {\em In--degree}, $k_{in}$, being the
number of {\em incoming edges} of a vertex and {\em Out--degree}, $k_{out}$,
being the number of its {\em outgoing edges}. In the following we will be
interested only in the outgoing degree, from here on simply referred to as to degree.}:
the {\em total number of its
connections} or its nearest neighbors. Because we are dealing with a weighted
network we can also introduce the {\em weighted node degree}, also
called {\em node strength}~\cite{BBPV2004}, namely
$s_i(t)=\sum_{j}\alpha_{ij}^t/(N-1)$. Its mean value averaged over the whole
network reads:
\begin{equation}
\label{eq:strength}
<s>(t) = \frac{1}{N}\sum_{i=1}^N s_i(t)\, .
\end{equation}
Let us observe that the normalization factor
$N-1$ holds for a population of
$N$ agents, self-interaction being disregarded, $<s>$ belongs hence to the interval $[0,1]$ and having eliminated the relic of the
population size, one can properly compare quantities calculated for networks made of different number of agents.
All these quantities evolve in time because of the dynamics of the opinions
and/or affinities. Passing to continuous time and
using the second relation of~\eqref{eq:themodel}, we obtain:
\begin{equation}
\label{eq:meandegevolv}
\frac{d}{dt}<s> = \frac{1}{N(N-1)}\sum_{i,j=1}^N \frac{d}{dt}\alpha_{ij}^t\, .
\end{equation}
Let us observe that the evolution of affinity and opinion can be
decoupled when $\Delta O_c=1$. For $\Delta O_c <1$, this is not formally
true. However on can argue for an
approximated strategy~\cite{opi3}, by replacing the step function $\Gamma_2$ by its time
average counterpart $\gamma_2$, where the dependence in $\Delta O^t_{ij}$ has been silenced.
In this way, we obtain form~\eqref{eq:meandegevolv}
\begin{equation}
\label{eq:meanalphaevolv}
\frac{d}{dt}<s> =\frac{\gamma_2}{N(N-1)}\sum_{i,j=1}^N \alpha^t_{ij}(1 -
\alpha^t_{ij})=\gamma_2 (<s>-<s^2>)\, ,
\end{equation}
where $<s^2> = \sum \alpha_{ij}^2/(N(N-1))$. Let us observe that $\gamma_2$ is
of the order of $1/N^2$ times, a factor taking care of the asynchronous
dynamics~\cite{opi3}.
In~\cite{evoden} authors proved that~\eqref{eq:meanalphaevolv} can be
analytically solved once we provide the initial distribution of node strengths
(see~\ref{app:analsol} for a short discussion of the involved
methods). Assuming $s_i(0)$ to be uniformly distributed in $[0,1/2]$, we get
the following solution (see Fig.~\ref{fig:alphamed}):
\begin{equation}
\label{eq:sanal}
<s>(t)=\frac{e^{\gamma_2 t}}{e^{\gamma_2 t}-1}-\frac{2e^{\gamma_2
t}}{(e^{\gamma_2 t}-1)^2}\log\left(\frac{e^{\gamma_2
t}+1}{2}\right)\, ,
\end{equation}
Using similar ideas we can prove~\cite{evoden} that the variance
$\sigma^2_{s}(t)=<s^2>-<s>^2$ is analytically given by
\begin{equation}
\label{eq:sigmaUD}
\sigma^2(t)=\frac{2e^{2\gamma_2 t}}{(e^{\gamma_2 t}-1)^2(e^{\gamma_2
t}+1)}-\frac{4e^{2\gamma_2 t}}{(e^{\gamma_2
t}-1)^4}\left[\log\left(\frac{e^{\gamma_2
t}+1}{2}\right)\right]^2\, .
\end{equation}
\begin{figure}[th]
\centerline{\psfig{file=AlphaEvolv21022008.eps,width=8cm}}
\vspace*{8pt}
\caption{Evolution of $<s>(t)$. Dashed line (blue on-line) refers to
numerical simulations with parameters $\alpha_c=0.5$, $\Delta O_c=0.5$
and $\sigma=0.3$. The full line (black on-line) is the analytical
solution~\eqref{eq:sanal} with a best fitted parameter $\gamma_2=1.6 \,
10^{-4}$. The dot denotes the convergence time in the opinion space
to the consensus state, for the used parameters affinities did not yet converge.
Let us observe in fact that affinities and opinions do converge
on different time scale~\cite{opi3}.}
\label{fig:alphamed}
\end{figure}
The comparison between analytical and numerical profiles is enclosed in Fig.~\ref{fig:alphamed}, where the evolution of $<s>(t)$ is traced. Let us observe that here $\gamma_2$ will serve as a fitting parameter, when testing the adequacy of the proposed
analytical curves versus direct simulations, instead of using its computed numerical value~\cite{opi3}.
The qualitative correspondence is rather satisfying, so confirming the correctness of the analytical results reported above.
Assume $T_c$ to label the time needed for the consensus to be reached. Clearly, $T_c$ depends on the size of the
simulated system~\footnote{In~\cite{pre,epl} it was shown that $T_c$ scales
faster than linearly but slower than
quadratically with respect to the population size $N$.}. From the above relation (\ref{eq:sanal}),
the average node strength at convergence as an implicit function of the population size
$N$ reads:
\begin{equation}
\label{eq:sTc}
<s>(T_c(N))=\frac{e^{\gamma_2(N) T_c(N)}}{e^{\gamma_2(N)
T_c(N)}-1}-\frac{2e^{\gamma_2(N) T_c(N)}}{(e^{\gamma_2(N)
T_c(N)}-1)^2}\log\left(\frac{e^{\gamma_2(N) T_c(N)}+1}{2}\right)\, ,
\end{equation}
where we emphasized the dependence of $\gamma_2$ and $T_c$ on $N$. However,
as already observed $\gamma_2(N)=\mathcal{O}\left(N^{-2}\right)$ and $T_c(N)=
\mathcal{O}\left(N^{a}\right)$, with $a\in (1,2)$. Hence
$\gamma_2(N)T_c(N)\rightarrow~0$ when $N\rightarrow \infty$ and thus
$<s>(T_c(N))$ is predicted to be a decreasing function of the population size $N$, which
converges to the asymptotic value $1/4$, a value identical to the initial average node
strength (see Fig.~\ref{fig:Ngradomedio}), given the selected initial condition. In
sociological terms this means that even when consensus is achieved the larger
the group the smaller, on average, the number of local acquaintances. This is a second
conclusion that one can reach on the basis of the above analytical developments.
\begin{figure}[th]
\centerline{\psfig{file=Alphac.eps,width=8cm}}
\vspace*{8pt}
\caption{Average node strength at convergence as a function of the population
size. Parameters are $\Delta O_c=0.5$, $\sigma=0.5$ and four values of
$\alpha_c$ have been used : ($\Diamond$) $\alpha_c=0$, ($\triangle$)
$\alpha_c=0.25$, ($\Box$) $\alpha_c=0.5$ and ($\bigcirc$)
$\alpha_c=0.75$. Vertical bars are standard deviations computed over $10$
replicas of the numerical simulation using the same initial conditions.}
\label{fig:Ngradomedio}
\end{figure}
\subsection{Small world}
\label{sec:smallworld}
Several social networks exhibit the remarkable property that one can reach an arbitrary far member of the community,
via a relatively small number of intermediate acquaintances. This holds true irrespectively of the size of the underlying network.
Experiments~\cite{milgram} have been devised to quantify the
\lq\lq degree of separation\rq\rq in real system,
and such phenomenon is nowadays termed the \lq\lq small world\rq\rq effect, also referred
to as the \lq\lq six degree of separation\rq\rq.
On the other hand several, models have been proposed~\cite{WS1998,NW1999} to construct complex
networks with the
small world property. Mathematically, one requires that the average
shortest path grows at most logarithmic with respect to the network size, while the network still displays a
large clustering coefficient. Namely, the network has an average shortest path
comparable
to that of a random network, with the same number of nodes and links, while
the clustering coefficient is instead significantly larger.
In this section we present numerical results aimed at describing the
time evolution of both the average shortest path and the clustering coefficient of
the social network emerging from the model. As before, the parameters are set so to induce the convergence to a
consensus state in the opinion space.
We will be particularly interested in their asymptotic solutions, terming the associated values respectively
$<\ell>(T_c)$ and $C(T_c)$ once the consensus state has been achieved.
In Fig.~\ref{fig:NevolCCrndEll} we report these quantities
(normalized to the homologous values estimated for a random network with identical
number of nodes and links) versus
the system size. The (normalized) clustering coefficient is sensibly larger than one, this effect being more pronounced the
smaller the value of $\alpha_c$. On the other hand the (normalized) average
shortest path is always very close to $1$.
Based on the above we are hence brought to conclude that the
social network emerging from the opinion exchanges, has the small world
property. This is a remarkable feature because the social
network evolves guided by the opinions, as it does in reality, and not result from
an artificially imposed recipe.
\begin{figure}[th]
\centerline{\psfig{file=fig17left.eps,width=7cm}
\quad \psfig{file=fig18left.eps,width=7cm}}
\vspace*{8pt}
\caption{Normalized clustering coefficient (left panel) and normalized
average mean path (right panel) as functions of the network size at the convergence
time. Parameters
are $\Delta O_c=0.5$, $\sigma=0.5$ and four values of
$\alpha_c$ have been used : ($\Diamond$) $\alpha_c=0$, ($\triangle$)
$\alpha_c=0.25$, ($\Box$) $\alpha_c=0.5$ and ($\bigcirc$)
$\alpha_c=0.75$. Vertical bars are standard deviations computed over $10$
repetitions.}
\label{fig:NevolCCrndEll}
\end{figure}
\subsection{Weak ties}
\label{ssec:weakties}
Social networks are characterized by the presence of hierarchies of well tied
small groups of acquaintances, that are possibly linked to other such groups via
\lq\lq weak ties\rq\rq. According to
Granovetter~\cite{Granovetter1983} these weak links are fundamental for the
cohesion of the society, being at the basis of the social tissue, so motivating the
statement \lq\lq the strength of weak ties\rq\rq.
The smallest group in a social network is composed by three individuals
sharing high mutual affinities, in term of network theory they form a {\em
clique}~\cite{AB2002}, i.e. a maximal complete graph composed by three
nodes. This can of
course
be generalized to larger maximal complete graphs, defining thus $m$-cliques.
The
degree of cliqueness of a social network is hence a measure of its
cohesion/fragmentation: the presence of a large number of $m$-cliques together
with very few, $m^{\prime}$-cliques, for $m^{\prime}>m$, means
that the population is actually fragmented into small pieces, of size $m$ not
strongly interacting each other.
We are interested in studying such phenomenon within the social network
emerging from the opinion dynamics model here considered, still operating in
consensus regime. To this end we proceed as follows. We introduce a
cut--off parameter $\alpha_f$ used to binarize the affinity matrix, which
hence transforms into a
an adjacency matrix $a$. More precisely, agents $i$ and $j$ will be
connected, i.e. $a_{ij}=1$, if and only if $\alpha_{ij}\geq \alpha_f$. Once
the adjacency matrix is being constructed, we compute the number of $m$--cliques in the
network. Let us observe that this last step is highly time consuming, being the
clique problem NP-complete. We thus restrict our analysis to the cases $m\in\{ 3,4,5 \}$.
For small values of $\alpha_f$ the network is almost complete, while for
large ones it can in principle fragment into a vast number of finite small groups
of agents. As reported in the inset of right panel of
Fig.~\ref{fig:cliques}, for $\alpha_f\sim 1$ only $3$--cliques are present.
Their number rapidly increases as $\alpha_f$ is lowered. On the other hand
for $\alpha_f\sim 0.98$ few $4$--cliques emerge while $5$--cliques appear
around $\alpha_f\sim 0.73$. This means that the social networks is mainly
composed by $3$--cliques, i.e. agents sharing high mutual affinities, that are
connected together to form larger cliques,
for instance $4$ and $5$--cliques by weaker links, i.e. whose mutual affinities are lower than the above ones.
Results reported in left panel of Fig.~\ref{fig:cliques} show that for specific
parameter values, still falling into the class deputed to the consensus dynamics, the model does
not present the weak ties phenomenon: $3$, $4$ and $5$-cliques are all
present at the same time for large values of $\alpha_f$. This is an important point that will
deserve future investigations. Let us observe here that the observed differences
stem from the social temperature.
\begin{figure}[th]
\centerline{\psfig{file=Simulaz20080626_003.eps,width=7cm}
\quad \psfig{file=Simulaz20080626_003.eps,width=7cm}}
\vspace*{8pt}
\caption{Number of $3$, $4$ and $5$--cliques in the social network once
consensus has been achieved. Parameters are $N=100$, $\Delta O_c=0.5$,
$\alpha_c=0.5$. Right panel, $\sigma=0.5$, the network exhibits the weak ties
property. Left panel, $\sigma=0.1$, the network does not display the weak ties phenomenon.}
\label{fig:cliques}
\end{figure}
\section{Conclusion}
\label{sec:conc}
Social system and opinion dynamics models are intensively investigated within simplified mathematical schemes.
One of such model is here revisited and analyzed. The evolution of the underlying network of connections, here emblematized
by the mutual affinity score, is in particular studied. This is a dynamical quantity which adjusts all along the system evolution, as follows a
complex coupling with the opinion variables. In other words, the embedding social structure is adaptively created and not a
priori assigned, as it is customarily done. Starting from this setting, the model is solved analytically, under specific approximations. The
functional dependence on time of the networks mean characteristics are consequently elucidated. The obtained solutions correlate
with direct simulations, returning a satisfying agreement. Moreover, the structure of the social network is numerically monitored,
via a set of classical indicators. Small world effects, as well weak ties connections, are found as an emerging property of the model. It
is remarkable that such properties, ubiquitous in nature, are spontaneously generated within a simple scenario which accounts for a minimal number
of ingredients, in the context of a genuine self-consistent formulation.
|
\section{Introduction}
\bigskip
\bigskip
\label{1intr1} Easter Island history is a very famous example of an evolved
human society that collapsed for over exploiting its fundamental resources
\cite{brander,dalton,anderies} that in this case were essentially in palm
trees. It were covering the island\cite{flenley} when, few dozens of
individuals, first landed around 400 A.D. Its advanced culture was developed
in a period of one thousand years approximately. Its ceremonial rituals and
associated construction were demanding more and more natural resources
especially palm trees. The over exploitation of this kind of tree, very
necessary as a primary resource (tools construction, cooking, erosion
barrier, etc.) was related with the collapse. In this paper, a mathematical
model concerning growing and collapse of this society is presented.
Different to usual like Lotka-Volterra models\cite{has,mur} where the
carrying capacity variation becomes from external natural forces, in this
work it is directly connected with the population dynamics. Namely,
population and carrying capacity are interacting dynamics variables, so
generalizing the Leslie model prey-predator.
The general mathematical treatment of a model describing such a
complex society is a very hard task and probably not unique. Our
aim is to settle the most simple model describing with acceptable
precision the evolution of Easterner society. With the idea of
writing a model that could be generalized to a more complex
system, we first divide the elements into two categories: the
\textit{resource} quantity $R_{i}$ with $i=1,2\cdots k$ and the
\textit{inhabitants} numbers (species) $N_{i}$ with $i=1,2\cdots
m$. With the concept of resources we are meaning resources in a
very large sense, it could be oil, trees, food and so on. The
several kind of resources are described by the index $i$. In
similar way, with the concept of inhabitants, we are meaning
different species of animals or internal subdivision of human
being in country or town or even tribes. Leaving the idea of a
constant quantity of resources that leads to the logistic equation
for the number on inhabitants\cite{logistic}, the aim of this
paper is to include in the dynamical description of the time
evolution of the system the resources too which can not be
considered as constant. A generalization of the logistic equation
to an arbitrary number of homogeneous species interacting among
the individuals (with non constant resources) can be written as
\begin{equation}
\frac{d}{dt}N_{i}=r_{i}N_{i}\left[ 1-\frac{N_{i}}{N_{ci}\left( R_{1}\cdots
R_{k}\right) }\right] -\sum\limits_{j=1\,,i\neq j}^{m}\chi _{ij}N_{j}N_{i}.
\label{log-inter}
\end{equation}%
Where $r_{i}$ is the usual growing rate for species $i$. In the denominator
it appears the carrying capacity of the system with respect to the number of
inhabitants $N_{ci}\left( R_{1}\cdots R_{k}\right) $. Beside the dependence
on the resources $R_{i}$ we could have also a dependence on an other species
that would be then a \textquotedblright resource\textquotedblright\ for some
other species. This fact is expressed even by the quantities $\chi _{ij}$
that in general are not symmetric expression ($\chi _{ij}\neq \chi _{ji}$)
since that the prey is a resource for the predator and not the inverse.
Similarly, for the resources we have:
\begin{equation}
\frac{d}{dt}R_{i}=r_{i}^{\prime }R_{i}\left[ 1-\frac{R_{i}}{R_{ci}}\right]
-\sum\limits_{j=1}^{m}\alpha _{ij}N_{j}R_{i}. \label{res-inter}
\end{equation}%
It is clear that set of equations (\ref{res-inter}) could be formally
included into Eq. (\ref{log-inter}) redefining the quantities $\alpha _{ij}$%
. Nevertheless, we shall keep this distinction for the sake of
clarity especially referred to resources, such as trees, oil or
oxygen, where the carrying capacity is not determined by other
species and can be considered a constant ($R_{ci}$). Also the
meaning of the parameters $r_{i}^{\prime }$ is more or less the
same of the analogous parameters $r_{i}$. It is suitable to define
$r_{i}^{\prime }$ as \textquotedblright renewability
ratios\textquotedblright , since describe the capacity of the
resources to renew itself and clearly are depending on the kind of
resource. For example, the renewability of the oil is clearly zero
since the period of time to get oil from a natural process is of
the order of geological time-scale processes. In general all
parameters of Eqs. (\ref{log-inter}) and (\ref{res-inter}) are
time dependent, including stochasticity. We can assume reasonably
slowly time-varying for the ancient societies so that we can
consider it as constant\cite{flores}, particularly the $\alpha
$'s. Anyway the set of $\alpha _{ij}$ is worthy of a more detailed
discussion. We can call this set of parameters
\textit{technological parameters} in the sense that they carry the
information about the capacity to exploit the resources of the
habitat. We shall see in the next section that the technological
parameters combined with the renewability ratios will be the key
point to decide wether, or not, a society is destined to collapse.
\section{Easter island collapse model}
\label{2east2} The particular history of Easter Island society presents
several advantages for modelling its evolution \cite{bologna}. In fact it
can be with very good approximation considered a closed system. The peculiar
style of life and culture allows us to consider a basic model where trees
are essentially the only kind of resources. Many of activities of the
ancient inhabitants involved the trees, from building and transport the
enormous Moai, to build boats for fishing, etc. In fact the cold water was
not adapt to the fish life and the impervious shape of the coast made
difficult fishing. Finally from historical reports it can be inferred that
the inhabitants did not change the way to exploit their main resource, even
very near to exhaust it so that we can consider the technological parameter
as constant. Considering Eqs. (\ref{log-inter}) and (\ref{res-inter}) for
one inhabitant species and one kind of resource, we obtain:
\begin{equation}
\frac{d}{dt}N=rN\left[ 1-\frac{N}{N_{c}\left( R\right) }\right] ,
\label{log-inter1}
\end{equation}
\begin{equation}
\frac{d}{dt}R=r^{\prime }R\left[ 1-\frac{R}{R_{c}}\right] -\alpha NR,
\label{res-inter1}
\end{equation}%
where we introduce the notation: $\alpha _{11}\equiv \alpha $. The unknown
function $N_{c}\left( R\right) $ has to satisfy few properties. For a
quantity of primary unlimited resource, $R\rightarrow \infty $, even $%
N_{c}\left( R\right) \rightarrow \infty $ that means that the
population can grow unlimited too. In the opposite case
$R\rightarrow 0$ clearly also the population must vanish,
$N_{c}\left( R\right) \rightarrow 0$, and finally when the
resource is constant we are back to the ordinary equation
(logistic) so that $N_{c}\left( R\right) = \textrm{ const}$. It is
clear that the choice of this relation is quite arbitrary but
following the simplicity criteria we can select $N_{c}\left(
R\right) =\beta R$, where $\beta $ is a positive parameter. This
choice formalizes the intuition that the maximum number of
individuals tolerated by a niche is proportional to the quantity
of resources. We note that in (\ref{res-inter1}) the interacting
term depend
on the variable $R$. Namely, for $R=0$ no variation of resource exist ($%
\frac{d}{dt}R=0$) corresponding to a biological criterion and different from
this one of reference\cite{basener}. A more sophisticate model should
include also fishing as resource and consider for $N_{c}\left( R\right) $ an
expression such as $N_{c}\left( R\right) =\beta R+R_{f}$ where $R_{f}$ is
the fishing carrying capacity. This resource was limited near the coast and
could not be fully exploited without boats, so that, we are going to neglect
this resource. We can rewrite Eqs. (\ref{log-inter1}) and (\ref{res-inter1})
as:
\begin{equation}
\frac{d}{dt}N=rN\left[ 1-\frac{N}{\beta R}\right] , \label{log-inter2}
\end{equation}
\begin{equation}
\frac{d}{dt}R=r^{\prime }R\left[ 1-\frac{R}{R_{c}}-\alpha
_{E}N\right] ,\,\,\,\,\textrm{where }\,\,\alpha _{E}\equiv
\frac{\alpha }{r^{\prime }}. \label{res-inter2}
\end{equation}%
The dimensionless parameter $\alpha _{E}$ is the ratio between the
technological parameter $\alpha $, representing the capability of to exploit
the resources, and $r^{\prime }$ the renewability parameter representing the
capability of the resources to regenerate. We will call $\alpha _{E}$
deforestation parameter since it gives a measure of the rapidity with which
the resources are going to exhaust and then a measure of the reversibility
or irreversibility of the collapse. Using the historical data we can have an
estimation of the parameters. At the origin ($t=0$) we can assume that the
trees were covering the entire island surfaces of 160 km$^{2}$. When the
first humans arrived to the island, around the 400 A.D., their number was of
the order of few dozens of individuals and it grew until to reach the
maximum $N_{M}\sim 10000$ around the 1300 A.D.
Finding the equilibrium points of Eqs. (\ref{log-inter2}) and (\ref%
{res-inter2}) we obtain (see section III, for stability):
\begin{equation}
N_{0}=0,\,\,\,\,R_{0}=R_{c}\,\,\,\,(\textrm{unstable-saddle-point})
\label{equin}
\end{equation}
\begin{equation}
N_{e}=\frac{\beta R_{c}}{1+\alpha _{E}\beta R_{c}},\,\,\,\,R_{e}=\frac{R_{c}%
}{1+\alpha _{E}\beta R_{c}}\,\,\,\,(\textrm{stable}).
\label{equir}
\end{equation}%
While the point $(N_{0},R_{0})$ of Eq. (\ref{equin}) represents the trivial
fact that in absence of human being the number of trees is constant
(carrying capacity). Eq. (\ref{equir}) describes the fact that, due to the
interaction humans-environment, the more interesting equilibrium point $%
(N_{e},R_{e})$ does not coincide with $(N_{c},R_{c})$ with $N_{c}=\beta
R_{c} $ since $\alpha \neq 0$. To study the stability of the point $%
(N_{e},R_{e})$, we have linearize the system of Eqs. (\ref{log-inter2}) and (%
\ref{res-inter2})around the equilibrium point. In fact, in the next section
we shall show that it is a stable equilibrium point.
\section{Equilibrium points and stability}
\label{equilibrium}
Let us first cast Eqs. (\ref{log-inter2}) and (\ref{res-inter2})
in term of dimensionless quantities; setting $\nu (t)=N/N_{c}$,
$\varrho (t)=R/R_{c}$ and $\tau =rt$, we have:
\begin{equation}
\frac{d}{d\tau }\nu =\nu \left[ 1-\frac{\nu }{\varrho }\right] \label{adnu}
\end{equation}
\begin{equation}
\frac{d}{d\tau }\varrho =\bar{r}\varrho \left[ 1-\varrho -\bar{\alpha}\nu %
\right] ,\,\,\,\,\bar{\alpha}\equiv \alpha _{E}N_{c},\,\,\,\,\bar{r}\equiv
\frac{r^{\prime }}{r}. \label{adro}
\end{equation}
For sake of clarity we rewrite also the equilibrium point:
\begin{equation}
\nu _{e}=\frac{1}{1+\bar{\alpha}},\,\,\,\,\varrho _{e}=\frac{1}{1+\bar{\alpha%
}} \label{equirad}
\end{equation}
with obvious meaning of the symbols. Perturbing the equilibrium point $\nu
(\tau )=\nu _{e}\left[ 1+\eta (\tau )\right] $ and $\varrho (\tau )=\varrho
_{e}\left[ 1+\varepsilon (\tau )\right] $ with $\eta (\tau )$ and $%
\varepsilon (\tau )$ infinitesimal functions, after straightforward algebra
we obtain Eq. (\ref{equir}):
\begin{equation}
\frac{d}{d\tau }\eta =-\eta +\varepsilon \label{lineta1}
\end{equation}
\begin{equation}
\frac{d}{d\tau }\varepsilon =-\frac{\bar{r}}{1+\bar{\alpha}}\left[ \bar{%
\alpha}\eta +\varepsilon \right] \label{lineps1}
\end{equation}
The eigenvalues of the system are:
\begin{equation} \label{eig}
\lambda_{1,2}=\frac{-(1+\bar{\alpha}+ \bar{r})\pm\sqrt{\left(1+\bar{\alpha}+%
\bar{r}\right)^{2} -4\bar{r}(1+\bar{\alpha})^{2}}}{2(1+\bar{\alpha})}
\end{equation}
Restricting ourself to the case of positive values of $\bar{\alpha}$, Eq. (%
\ref{eig}) shows that both the eigenvalues always have a real negative part,
so that the equilibrium point $(\nu_e,\varrho_e)$ is a stable equilibrium
point. More in detail we have that for
\[
\bar{r}\leq \frac{1}{4},\,\,\bar{\alpha}\geq 0\,\,\,\,\textrm{and for}\,\,\,\,%
\bar{r}>\frac{1}{4},\,\,\bar{\alpha}\leq \frac{\left( \sqrt{\bar{r}}%
-1\right) ^{2}}{2\sqrt{\bar{r}}-1}
\]
the eigenvalues are real and negative so that the equilibrium point is
reached in exponential damped way, otherwise the eigenvalues acquire an
imaginary part and the system reach the equilibrium point via exponential
damped oscillations.
\begin{figure}[h]
\par
\begin{center}
\epsfig{file=pasqua.eps,height=6cm,width=12cm,angle=0}
\end{center}
\caption{\label{figlor} $N(t)/N_c$ versus time. In this numerical
example $N(0)/N_c\sim 0.8$ and $N_{Max}/N_c\sim 3$}
\end{figure}
\section{Collapse condition}
Even if, mathematically speaking, the stable point $(N_{e},R_{e})$
is an acceptable result we have to take in account the biological
constraints that allow to a specie to survive. A reasonable number
of individuals is required for viability of a given species
\cite{beg,bel}. This is so because genetic diversity, social
structures, encounters, etc., need a minimum numbers of
individuals since under this critical numbers the species is not
viable and collapse. It is worthy to stress that while the trees
can reach an equilibrium point without the humans, Eq.
(\ref{equin}), the opposite does not hold, as stated by Eq.
(\ref{equir}).
Calling the minimum number of humans $N_{min}$ we can find a upper bound for
the parameter $\alpha _{E}$ so that a civilization can survive. Imposing the
condition that at the equilibrium point $N_{e}\geq N_{min}$ we obtain:
\begin{equation}
\alpha _{E}\geq \frac{1}{N_{min}}-\frac{1}{\beta R_{c}}\textrm{
(collapse condition)}. \label{bond}
\end{equation}%
As further simplification of inequality (\ref{bond}), we assume that $%
N_{min}\ll \beta R_{c}$ and we find that $\alpha _{E}N_{min}\geq 1$ or $%
\alpha N_{min}\geq r^{\prime }$. It is a natural condition since it tells
that collapse exists when the production rate $r^{\prime }$ is minor than
the deforestation rate $\alpha N_{min}$.
More in general it can be showed, numerically, that considering
the standard case with the starting population number
$N(0)<N_{c}=\beta R_{c}$, we can have a solution that can exceed
the value $\beta R_{c}$ (depending on initial conditions $N(0)$
and $\frac{dN}{dt}|_{t=0}$). It is worthy to stress that in the
case of ordinary logistic map the population number never can
exceed this limit value\cite{logistic}. In the example of Fig.
\ref{figlor}, the maximum of $N$ is reached at a value that is
almost three times $\beta R_{c}$. Then, according to the region of
the parameters that we are considering, the paths to the final
equilibrium is exponentially fast, reaching eventually the point
$N_{min}$ and collapsing.
\section{Deforestation rate estimation}
As we saw in the previous section, the collapse condition (15) gives a
sufficient condition on the deforestation rate per individual $\alpha$ . On
other hand, the last period of tree extinction was governed essentially by
the deforestation rate. In this way, we have the rate of tree extinction
\begin{equation}
\frac{1}{R}\frac{dR}{dt}\sim -\alpha N. \label{16}
\end{equation}%
As discussed, the path to the equilibrium point is exponentially fast, so
that a rough estimation of the left side of Eq. (\ref{16}) is the time scale
of the deforestation, $\tau _{F}$, while the right side can be taken at the
end of the collapse process (the equilibrium point):
\begin{equation}
\frac{1}{\tau _{F}}\sim \alpha N_{F}, \label{17}
\end{equation}%
being $N_{F}$ the final number of individuals. It can be deduced \cite%
{pointing} that the range of is $\tau _{F}\sim 100$ yrs. to $\tau _{F}\sim
300$ yrs. and $N_{F}\sim 3000$, the rate of deforestation (per individual)
could be estimated as:
\begin{equation}
\alpha \sim \frac{1}{\tau _{F}N_{F}}\textrm{ ( yrs.
individual)}^{-1}
\end{equation}%
giving a range
\begin{equation}
1.1\times 10^{-6}<\alpha <3.3\times 10^{-6}\textrm{ ( yrs.
individual)}^{-1}.
\end{equation}%
This estimation has validity in the case of exponential decay which is the
our case, as it has been showed by the analysis performed in Sec. \ref%
{equilibrium}. Assuming the number of trees as proportional to the area $A$
we can now estimate the rate-deforestation-area. The island has a surface at
order of $A\sim 160$ km$^{2}$ and initially it can be supposed that was
covered of trees so that we can estimate:
\begin{equation}
0.5<\frac{dA}{dt}\sim \frac{A_{0}}{\tau _{F}}<1.6\,\,(\textrm{km}^{2}/\textrm{yrs%
}).
\end{equation}%
As comparison we can consider that in the last 500 years the
deforestation of amazonian forest rate is 15
$(\textrm{km}^{2}/\textrm{yrs})$. Considering that in 500 years
the deforestation technology became more and more efficient,
especially in the last century, we can consider it as an upper
limit, giving us an idea of the technological change. In the human
history there are several examples of over exploiting the natural
resources even if not so known as Easter island. In particular, the Cop\'{a}%
n Maya history\cite{wil} has certain similarity with respect to the
technology level and the over exploiting of the natural resources. In short,
this ancient civilization reached almost 20.000 individuals and declined to
5000 individuals in the 9th century. Using the estimation of the
technological parameter $\alpha $ obtained for eastern island civilization,
we get a collapse time from Eq. (\ref{17}) that is $60<\tau <180$ years. The
collapse time based on historical reports is $\tau \sim 100$, showing that
the adopted model is consistent with the available data. The estimation of
the parameter is consistent with the idea that similar civilizations, in
technological meaning, have similar capacity to exploit the natural
resources.
\section{Concluding remarks}
A mathematical model considering the interaction among carrying capacity and
population in an isolated system has been considered. The model takes in
account the fact that a population can over exploit the carrying capacity
without saturate, a fact of relevant importance on the path leading to a
collapse of a society. Its application to the collapse of the Easter Island
civilization has been presented. An estimation of the technological
parameter $\alpha $ is obtained and applied to an other ancient
civilization, the Cop\'{a}n-Maya, with a reasonably precise expectation
about their collapse time. All confirming the consistency of the adopted
model. On the other hand, its relative reasonable prediction suggests a
possible extension to more complex system. The effort to mathematical
modelling of ancient civilizations \cite{mur,flores,hor} could be important
considering the actual human growing and resources exploitations. An
adequate equilibrium between competition, demand and exploitation is the key
of surviving.
\label{4conc4}
\section*{Acknowledgments}
The authors acknowledge support from the project UTA-Mayor 4787 (2006-2007)
and CIHDE-project.
\bigskip
|
\section{Introduction}
KH~15D is a system with unique and dramatic photometric behavior that was discovered by \citet{kearns98} during a variability study of the extremely young cluster NGC~2264. Basic properties of the star, including its distance (760~pc) and classification as a weak-lined T~Tauri star (WTTS) were reported by \citet{hamilton01}. It is now known to be a young ($\sim$3~Myr) eccentric binary system embedded in a nearly edge-on circumbinary disk that is tilted, warped and precessing with respect to the binary orbit \citep{winn04, johnsonwinn04, chiang04, johnson04}. When first discovered in 1995, the disk was oriented relative to our line of sight in such a way that it covered most of the orbit of one star (now known as Star~B) but only a small portion of the orbit of the other (Star~A). During the past 14 years, precession of the disk has caused the sharp edge of the opaque screen to gradually cover all of the orbit of Star~B and most of the orbit of Star~A. As a result, we have not seen Star~B (except by reflected light) since 1995, and have seen direct light from Star~A for progressively less of its 48.4 day orbital cycle \citep{hamilton05, leduc09}. \citet{winn06} have produced a model of the system that accurately accounts for most of the photometric and radial velocity observations. They estimate the masses of Star~A and Star~B to be 0.6 and 0.7 $M_{\odot}$ and the radii to be 1.3 and 1.4 $R_{\odot}$, respectively. The spectral class of Star~A is K6/K7 \citep{hamilton01, agol04}.
The system's geometry and fortunate orientation have provided astronomers with a unique opportunity during the past decade and a half to study the structure of both the gas and the dust components of an evolved protoplanetary disk. \citet{herbst08} argued that the particulate grains making up the solid component of the disk, the ``grain disk", are probably of sand (millimeter) size or larger. It is possible that KH~15D possesses exactly the sort of settled particulate disk originally envisioned by \citet{safronov69} and \citet{goldreich73} as a prelude to the formation of planetesimals through gravitational instability \citep[see also][]{youdin08}. The gas disk at this stage is predicted to have a much larger scale height than the particulate disk since it is partially supported by pressure. Because Star~A of the KH~15D system changes elevations with respect to the edge of the grain disk, during the past decade we were presented with an opportunity to observe different path lengths through any accompanying gas disk for signs of absorption. We know that there is some gas still present in the disk because the star has an active magnetosphere and drives a weak bipolar jet \citep{hamilton03,tokunaga04,deming04,mundt09,hamilton09}. KH~15D has the potential to tell us about an important missing link in the planet formation process: the transition from protoplanetary disk to debris disk.
In this paper, we constrain properties of the gas component of the KH~15D circumbinary disk by carefully analyzing the \ion{Na}{1}~D absorption lines visible in the spectrum of Star~A while at various heights above the occulting disk. In \S\ref{observations} we describe the echelle spectra available for this analysis, their reduction, and the basic features of the observed \ion{Na}{1} absorption profile. In \S\ref{measuring} we discuss how \ion{Na}{1} column densities are computed. In \S\ref{ISM} we measure interstellar \ion{Na}{1} absorption toward KH~15D and discuss how this affects our absorption measurements. In \S\ref{scaleheight} we discuss the implications for the scale height of the gas disk, and in \S\ref{graindisk} we further discuss the interpretation of the data in terms of what it may reveal about the putative gas disk of KH~15D. Finally, in \S\ref{summary} we summarize the insight into the KH~15D disk revealed by this investigation.
\section{Observations} \label{observations}
\subsection{Spectral Data}
Over the past few years, a number of high-resolution echelle spectra of KH~15D have been obtained with Star~A at different elevations with respect to the grain disk edge. Some of these spectra include the \ion{Na}{1}~D doublet at 5895.9242 and 5889.9510 \AA, which allows us to search for evidence of absorption by neutral sodium gas which may be associated with the disk. These spectra were taken with the High Resolution Echelle Spectrometer \citep[HIRES;][]{vogt94} on Keck~I at the W.~M.\ Keck Observatory in Hawaii, the Ultraviolet and Visual Echelle Spectrograph \citep[UVES;][]{dekker00} on the Kueyen Telescope at the European Southern Observatory's Very Large Telescope (VLT) facility on Cerro Paranal in Chile, and the High Resolution Spectrograph \citep[HRS;][]{tull98} on the Hobby-Eberly Telescope \citep[HET;][]{ramsey98} at McDonald Observatory in Texas. A log of the spectra analyzed in this paper is given in Table~\ref{cds}.
The Keck/HIRES data were taken at a resolution $R \equiv \lambda/\Delta\lambda \sim 45,000$ in 2003 February and 2005 February (W.~Herbst, P.I.) The reader is referred to \citet{hamilton09} for a full desciption of the reduction procedures employed. The VLT/UVES data were taken at a resolution $R \sim 44,000$ in two separate observing runs, one in 2001 November (C.~Bailer-Jones, P.I.) and one in 2004 December (R.~Mundt, P.I.) Data were reduced using the UVES pipeline as described in \citet{hamilton03} and \citet{hamilton09}. The HET/HRS data were taken at a resolution $R \sim 30,000$ in 2005 December and 2006 February (W.~Herbst, P.I.) Data were reduced using IRAF \citep{IRAF} routines to subtract the bias, perform flat-field corrections, remove scattered light and cosmic rays, extract the echelle science and sky orders, and calibrate the wavelength solution using images of a Th-Ar hollow cathode taken the same night as the science data. IDL routines were written to subtract the sky spectra, normalize the science spectra, and convert to heliocentric velocities.
The comparison spectrum used here is of HD~36003, which has a spectral type of K5~V \citep{cenarro07}. The spectrum was taken on the night of 2003 February 8 with Keck/HIRES. This star has a parallax of 0.077\arcsec\ \citep{hipparcos}, making it close enough to alleviate worries about significant ISM contamination. While the spectral type is slightly earlier than that of the visible component of KH~15D (K6/K7), the match with other photospheric lines and with the extreme wings of the \ion{Na}{1}~D lines is excellent. The slightly later spectral class of KH~15D is offset by its lower surface gravity relative to a main sequence K6/K7 star.
We initially attempted to use other comparison stars that more closely match KH~15D's spectral type and evolutionary phase, but HD~36003 had the best overall fit to the spectra of KH~15D. Spectra of Hubble~4 and LkCa~7, which are nominally T~Tauri stars (TTS) of spectral class K7, were compared with KH~15D, but the fit with the clearly uncontaminated spectral features was significantly worse for both T~Tauri spectra than for HD~36003. This is most likely due to the larger rotational broadening of the features in these two TTS as compared to KH~15D, which has $v$~sin~$i$ = 6.9~$\pm$~0.3~km/s \citep{hamilton05}, and HD~36003, which has $v$~sin~$i$ $<$ 3~km/s \citep{luck06}. Hubble~4 has $v$~sin~$i$ = 14.6~$\pm$~1.7~km/s \citep{johnskrull04}, while LkCa~7 has $v$~sin~$i$ = 11~km/s \citep{glebocki01}. In addition, LkCa~7 has been noted to be a poor comparison star due to anomalous colors, possibly caused by a cooler companion \citep{gullbring98}. For these reasons, we chose to use HD~36003 as our comparison star.
In addition to this comparison star, the bright, nearby B2~III star HD~47887 was observed to place constraints on interstellar \ion{Na}{1} absorption near KH~15D (\S\ref{ISM}).
\subsection{The \ion{Na}{1}~D Profile of KH~15D} \label{profile}
Figure~\ref{compspec_sm} shows each of the analyzed KH~15D spectra in the region of the \ion{Na}{1}~D lines overlaid on the comparison spectrum, which has been appropriately shifted for each night of data. To accurately align the comparison star spectrum in radial velocity with the KH~15D spectra, we used a $\chi^2$ minimization technique based on the portion of the spectrum longward of the \ion{Na}{1}~D$_1$ line ($\sim$5900--5930~\AA), which contains several stellar absorption features due to \ion{Fe}{1} and \ion{Ti}{1}. Formally, the comparison star spectrum should be shifted by steps in log($\lambda$) \citep{tonry79}, but beause our analysis region includes such a small range of wavelengths ($\sim$15~\AA), the $\chi^2$ analysis is performed using uniform steps in $\lambda$. The part of the spectrum within the wings of the \ion{Na}{1}~D lines was excluded from the fitting, because we are attempting to measure very small discrepancies in the deepest part of these lines.
Visual inspection of the KH~15D spectra overplotted on the shifted comparison star spectrum clearly reveals that the two match well except at the very bottom of the \ion{Na}{1}~D lines. Figure~\ref{compspec_big} shows an enlarged view of the region around the \ion{Na}{1}~D lines to make the absorption and emission features more clearly visible. Here we find that the KH~15D \ion{Na}{1}~D line cores deviate from those of the comparison star spectrum, showing some potential filling on the blue side (though telluric emission makes this difficult to see in some panels of Figure~\ref{compspec_sm}) and excess absorption on the red side of the profile. Since this absorption feature appears in all 12 spectra, regardless of which telescope or reduction procedure was employed, we are confident that it is real. As this absorption is asymmetric and not seen in the comparison star spectrum, this additional sodium absorption is most likely caused by something other than the photosphere of Star~A. We turn now to an examination of the source of this absorption and, in particular, whether it could be associated with the gaseous component of the circumbinary disk.
\section{Analysis} \label{analysis}
\subsection{Measuring the \ion{Na}{1}~D Profile} \label{measuring}
Following the procedure of \citet{redfield07}, we began the process of quantifying the excess \ion{Na}{1} absorption in KH~15D by taking the ratio between each available spectrum and the shifted comparison star spectrum after the comparison star spectrum had been resampled to have the same pixel sizes as the KH~15D spectrum by rebinning. The VLT and HET spectra are both at lower resolutions than the Keck spectra, so the comparison spectrum was only resampled to larger pixel sizes. The results of this process are shown in Figures~\ref{ratiospec_sm} and~\ref{ratio_big} where we have effectively taken out the baseline stellar spectrum, leaving only the excess absorption and emission. We discuss each of these in turn in \S\S\ref{naemission} and \ref{quant}.
\subsubsection{\ion{Na}{1} Emission}\label{naemission}
While the dominant feature in these ratioed spectra is the red-shifted excess absorption feature clearly visible in the original spectra, this more detailed look also reveals a more subtle feature. All of the spectra show evidence of less absorption (or excess emission) at or close to the radial velocity of Star~A. This emission line is especially visible in the 2003 February Keck/HIRES data (see Figure~\ref{ratiospec_sm}) where the radial velocity of Star A is closest to the systemic velocity of 18.6~$\pm$~1.3~km/s \citep{winn06}, placing the emission feature close to the deepest part of the absorption feature in the ratioed spectra. Regardless of the source of this emission, in order to accurately measure any excess absorption, this emission feature must be taken into account in our model.
It is possible that the apparent excess emission is due to the spectral mismatch between the comparison star and KH~15D, but we believe that it is likely that Star~A actually does have an emission core within the stellar \ion{Na}{1}~D absorption features. Three arguments favor the view that the excess emission is real and associated with the star. We discuss each of these in turn.
First, we note that the emission feature is rather variable. On some nights the emission can be seen very strongly (e.g.\ 2005 December 22) and on others it is barely present (e.g.\ 2003 February 8). While we acknowledge the danger of over-interpreting ratioed spectra, especially deep in the cores of strong lines, the degree of variation does seem larger than can easily be accounted for by simply random noise. If the apparent emission arose only from a mismatch of the spectral types then it should be constant in time.
Another argument is that the emission closely matches the predicted radial velocity of Star~A in each spectrum, even as Star~A's radial velocity varies from -0.27--11.17~km/s during the course of our observations. This is just what is expected for emission by gas accreting onto the star. The predicted radial velocities of Star~A are marked in each frame of Figure~\ref{ratiospec_sm} for reference.
Finally, Star~A is known to have an active magnetosphere and bipolar jet \citep{hamilton03}. In this sense it is like a Classical T Tauri Star (CTTS), although the mass accretion rate is probably about an order of magnitude less than for a typical CTTS \citep{mundt09}. Emission in \ion{Na}{1} related to the active magnetospheres and jets has been observed in other T~Tauri stars, particularly when they are oriented such that we view them nearly edge-on to their disks \citep{appenzeller05}, as is the case here. Furthermore, in some spectra the emission lines appear extended to the blue. This is also a common feature of CTTS spectra \citep{appenzeller05} and it may be that we are detecting a rather weak version of the same phenomenon here.
On the whole, it appears reasonable to suppose that the emission lines seen at the radial velocity of Star~A in all of the spectra in Figure~\ref{ratiospec_sm} are, in fact, arising within the magnetosphere of Star~A, although the physical nature of the excess emission is not critical to the process of correcting for it in the calculation of the absorption column density.
We note that in some spectra (e.g.\ the Keck spectra from 2003 February) there is also a considerable amount of night sky emission which was not fully accounted for during the reduction process. Fortunately, in each case it is sufficiently shifted in radial velocity with respect to KH~15D that we have not needed to remove it from the line profiles because it does not affect the fitting results.
\subsubsection{Quantifying the Excess \ion{Na}{1} Absorption}\label{quant}
In order to most accurately quantify the redshifted excess absorption, we have found it necessary to take into account the variable emission that follows the stellar velocity. Since the \ion{Na}{1}~D line is a doublet we have two absorption and two emission features to model in each spectrum. We have experimented with a variety of different fitting techniques employing constrained and unconstrained Gaussian fits to the emission and absorption components. Given the noise level of the data and uncertainties about the physical source of the features we have, in the end, adopted a procedure quite similar to that employed by \citet{redfield07}. Its implementation here is described below.
For the two \ion{Na}{1}~D lines in the ratioed spectra, we fit a single Gaussian absorption component and a single Gaussian emission component to each. These four Gaussians (two emission and two absorption) are fit simultaneously, after being convolved with the appropriate line spread function for each instrument.
The two emission features are fixed at the predicted radial velocity of Star~A for each night of data, with the strength varying freely. However, the area contained within the Gaussian curve used to model the \ion{Na}{1}~D$_2$ emission line is fixed to be twice that of the Gaussian for the \ion{Na}{1}~D$_1$ line, due to the relative oscillator strengths of the two transitions \citep{morton03}, assuming optically thin \ion{Na}{1} gas.
The three important parameters for fitting the Gaussian absorption features are the column density, the radial velocity of the maximum absorption, and the Doppler width. Doppler width, also called the b-parameter or velocity-spread parameter, is the dispersion in velocities multiplied by $\sqrt{2}$. This measure of the width of the line is believed to be due to the Doppler effect resulting from turbulent and thermal motions \citep{spitzerbook}.
For the absorption fits, the equivalent widths of the two Gaussian features were also fixed to be 2:1, again corresponding to an assumption of optical thinness. For our initial fit, the radial velocities and widths of the two Gaussian absorption features were allowed to vary freely and were chosen by $\chi^2$ minimization, with the Doppler widths of the two absorption features fixed to be the same, and the radial velocities fixed to be identical. This yielded absorption features with an average radial velocity of 16.2~$\pm$~3.3~km/s, consistent with the systemic velocity of 18.6~$\pm$~1.3~km/s \citep{winn06}, and an average Doppler width of 5.6~$\pm$~2.5~km/s.
For our final absorption measurements, the two Gaussian emission features at the radial velocity of Star A and the two Gaussian absorption features were simultaneously fit to the observed profiles with the Doppler widths fixed to a width that produced the best fit by eye, only varying the column densities (fixed 2:1) and radial velocities (fixed to be identical) of the lines for the $\chi^2$ minimization. This results in a measurement of the absorbing column density of \ion{Na}{1} necessary to create the observed features. The results of our fits are listed in Table~\ref{cds} and shown in Figure~\ref{columndensities}.
For internal consistency in our analysis when comparing spectra from different telescopes and instruments, error bars on the individual spectral datapoints were calculated using the standard deviation of a range of datapoints within the normalized continuum, so that every datapoint within a given spectrum is given the same error value. The error bars on the column densities result from the standard deviation between three methods of fitting: 1) freely varying all three parameters (radial velocity, Doppler width, and column density), 2) fixing the Doppler widths, and 3) fixing the radial velocity to match the systemic velocity. This is obviously only a crude estimate of the error bars, since they are estimated from only three measured values. However, propagation of errors results in rather similar error bars and the agreement among the various measurements on different nights gives added confidence that these error bars are reasonable.
One source of error here is the subtraction of telluric \ion{Na}{1}~D emission lines. In the spectra where the radial velocities of telluric emission lines are closer to the $\sim$18.6~km/s systemic velocity (e.g.\ the 2001 November spectrum), it is more difficult to measure the column densities. Another source of error is the widely varying strength of the stellar emission lines.
An additional source of systematic error may be our slightly imperfect comparison star. Due perhaps to the evolutionary state of KH~15D, as discussed in $\S\ref{measuring}$, an earlier spectral type main sequence star actually matches the spectrum better than an identical spectral type T~Tauri star spectrum, particularly in the deep \ion{Na}{1}~D lines. However, as noted above, using a non-identical comparison star could introduce a small systematic error in the derived \ion{Na}{1} column densities, as the \ion{Na}{1}~D lines are slightly different in the K5~V star \citep{tripicchio97}. Such a mismatch is a possible explanation for the apparent excess flux shortward of Star~A's central velocity, mentioned above as also perhaps arising in the magnetospheric emission of Star~A. To a certain extent, our fitting procedure should remove this effect no matter what its cause, but there may well be a residual small effect on the column densities that is systematic in nature.
\subsection{Correcting for Interstellar Absorption} \label{ISM}
Figures~\ref{compspec_sm} and~\ref{compspec_big} show that there is indeed excess \ion{Na}{1} absorption present for both of the \ion{Na}{1}~D lines. Additionally, the other absorption lines in the stellar spectrum match up very well with absorption lines in the spectrum of the comparison star. This alone is strong evidence that there is some absorbing sodium gas toward the KH~15D system. Whether or not this absorbing material is circumstellar needs further investigation, and in particular, consideration of whether the absorption could arise in the general interstellar medium (ISM) between us and the star.
The only measurement of interstellar sodium in the general direction of KH~15D that we could find in the literature was for S~Mon, a well known spectral type O7 member of NGC~2264 located approximately 25\arcmin\ from KH~15D and likely at about the same distance from Earth. \citet{welty94} measure an absorbing component at roughly the systemic velocity of KH~15D, but much weaker (log~$N_{\rm Na I}$ $\sim$~11.8~cm$^{-2}$), with even weaker absorption extending blueward to a radial velocity of about 0~km/s. Not only is this feature not strong enough to account for our observations, but the general shape does not match our observations. However, the distance between the sight line to this star and to KH~15D could result in significantly different interstellar absorption profiles.
To get a better constraint on any possible interstellar medium contribution to the \ion{Na}{1}~D feature in KH~15D, we observed the bright B2~III star HD~47887, a cluster member located only 37\arcsec\ away. A spectrum was obtained with Keck/HIRES on 2003 February 9 to search for evidence of \ion{Na}{1} absorption in that star. Figure~\ref{bstar} shows the \ion{Na}{1}~D profile of the star and, indeed, two components of interstellar \ion{Na}{1} are clearly visible. We have fit them using the same methods described in $\S\ref{analysis}$ above for KH~15D, shown in the figure. The narrower, deeper component has a column density of log $N_{\rm Na I}$ = 12.002~$\pm$~0.002 cm$^{-2}$, with a Doppler width of 1.35~$\pm$~0.01~km/s and a radial velocity of 6.24~$\pm$~0.05~km/s according to this fit. The wider, shallower feature has a log $N_{\rm Na I}$ = 12.093~$\pm$~0.001 cm$^{-2}$, with a Doppler width of 10.88~$\pm$~0.02~km/s and a radial velocity of 15.65~$\pm$~0.05~km/s.
This analysis shows that there may well be a significant contribution to the excess absorption observed in KH~15D from the general interstellar medium. However, it is unlikely that it can fully explain the excess absorption for the following reasons. First, the column densities of \ion{Na}{1} measured in the KH~15D spectra are, in most cases, significantly larger than the total column density measured for HD~47887. Second, only the broad shallow feature is likely to be contaminating our measurements because the narrow deep feature lies at a radial velocity close to that of Star~A in most of our spectra. Its influence would be to reduce the apparent amount of emission in the \ion{Na}{1} emission feature discussed in \S\ref{naemission}, but it cannot be masquerading as the obvious redshifted absorption component seen in Figure~\ref{compspec_sm} because of the radial velocity difference.
The broad, shallow feature, on the other hand, does nearly match the measured radial velocities of the excess absorption in the KH~15D spectra, but is not as deep. It is probably a blend of several weak components at slightly different radial velocities leading to its broad, shallow nature. If this absorption feature appears at exactly the same strength in the direction of KH~15D as it does in HD~47887 only 37$\arcsec$ away, then it is not strong enough to account for all of the absorption, although it does need to be corrected for in assessing how much \ion{Na}{1} gas can reasonably be associated with the KH~15D disk. Our measurements have an average log~$N_{\rm NaI}$ of 12.5~cm$^{-2}$. Subtracting the contribution of the ISM as measured in the spectrum of HD~47887 leaves log~$N_{\rm NaI} \sim$~12.3~cm$^{-2}$, which is our best estimate of the column density associated with the disk of KH~15D. Formally, the error bar is log~$N_{\rm NaI} \sim$~0.1~cm$^{-2}$, but we note that the possibility of systematic errors and of contributions to the column from gas local to NGC~2264 but not associated with the KH~15D disk means that our measurement may be conservatively viewed as a rather strict upper limit. Certainly the column density of \ion{Na}{1} associated with the putative gaseous disk of KH~15D cannot have a column density much in excess of our measurement of log~$N_{\rm NaI}$ $\sim$~12.3~cm$^{-2}$.
\section{Scale Height of the Gaseous Disk}
\subsection{Comparing our Measurements to Theory}\label{scaleheight}
Figure~\ref{columndensities} shows that we have not been able to detect a significant variation of \ion{Na}{1} column density as Star~A moved relative to the grain disk edge. This is unfortunate because such a detection would have been clear evidence that the excess \ion{Na}{1} absorption is arising within the gaseous component of the disk. We now inquire whether the failure to detect such variation is an argument that the excess absorption that we have detected could not, in fact, be associated with the disk.
The exact geometry of the KH~15D system is not known and, in particular, we do not know the location of the sharp edge caused by the particulate disk. Our line of sight may intercept the disk (it is perhaps more accurately referred to as a ring in this context) at its inner edge, outer edge, or somewhere between \citep[these possible disk geometries are illustrated very clearly in Figure~6 of][]{winn06}. The inner edge is possibly at the 3:1 resonance with the binary system \citep[$\sim$0.6~AU;][]{herbst08} and the outer edge is probably limited to 5~AU or less in order to precess \citep{chiang04}. Assuming that the particulate disk marks the mid-plane of the gaseous disk and that our line of sight is nearly along the disk plane, we may then equate the apparent distance of Star~A from the opaque disk edge ($\Delta$) with the height of the gas layer that our line of sight to Star~A is probing. Of course, this is a simplification because the disk must be warped and because we are not really looking exactly along it. In the absence of more detailed knowledge of the system's geometry, it is a reasonable approximation.
The scale height ($H$) of the gas layer may be estimated under the assumption of hydrostatic equilibrium in the direction perpendicular to the disk ($z$). Assuming an ideal gas at a temperature ($T$) and constant vertical component of the local gravity ($g_z$), one may write
$$H = {g_z^{-1} \biggr[{P \over \rho}\biggr]} = {g_z^{-1} \biggr[{kT \over {\mu m_H}}\biggr]}$$ where $k$ is the Boltzmann constant, $\mu$ is the mean molecular weight of the gas and $m_H$ is the mass of the hydrogen atom. We further assume that the disk gas has a cosmic abundance of helium, is composed of predominantly neutral atoms, and that the Na atoms are well mixed with the H and He. In this case, we may set $\mu = 1.3$.
$T$ is expected to be in the range of about 150 to 400~K, depending on the exact location of the edge. The local vertical component of gravity, $g_z$, may be set by either the self-gravity of the disk or by the vertical component of the stellar gravity. In the first case, $g_z= 2 \pi G \Sigma$, where $\Sigma$ is the mass per unit area in the local disk, and in the second case $g_z = \Omega^2 z$ for small $z$, where $\Omega$ is the angular velocity of the disk material \citep{goldreich73}. For us, the more interesting limit is the weakest gravitational field that could be present, so we ignore any possible self-gravity. The range of motion of Star~A in our data set is only about 3 stellar radii ($R_A$, where $R_A$~= 1.3~$R_\odot$), so we take $z$~=~1 (in units of $R_A$) for this approximate calculation.
With the assumptions above, we estimate that a putative gas disk in hydrostatic equilibrium would have a measured scale height of $H$~= 14~(2600)~$R_A$ at the inner (outer) edge of the disk, located at 0.6~(5)~AU. These scale heights are shown in Figure~\ref{columndensities}. This is too large a scale height for us to have measured, given the limited range in $\Delta$ probed by Star~A during our observations and the precision of our measurements. Conversely, one can argue that if the the slightly decreasing trend of \ion{Na}{1} column density with $\Delta$, visible in Figure~\ref{columndensities}, is real, then the occulting edge of the gas disk is likely to be the inner edge of the disk and there may well be a contribution to g$_z$ from the self-gravity of the disk (i.e.\ it would need to be a few times larger than the stars alone could provide).
\subsection{The Settled Grain Disk of KH~15D} \label{graindisk}
We did not detect any statistically significant dependence of the excess \ion{Na}{1} absorption on the height of Star~A above the occulting edge. One expects from the order-of-magnitude analysis in \S\ref{scaleheight} that the scale height will be much greater than the few stellar radii range of $\Delta$ for which we have data. By contrast, the scale height of the grain disk responsible for the extinction must be less than 1~$R_A$. It has been represented in models based on the light curve as a sharp ``knife edge" \citep{herbst02,winn06}, and may be estimated from Figure~1 of the Supplement of \citet{herbst08} as having the value 0.7~$R_A$. While there is uncertainty in all of these estimates it seems clear that the scale height of the particulate disk is at least one order of magnitude less than the scale height of any gaseous disk associated with it.
Furthermore, our lack of detection of a strong \ion{Na}{1} feature associated with the steep increase in extinction at the grain disk boundary demonstrates again how different (evolved) this obscuring matter is from interstellar grains. The extinction in the I~band for our most obscured data point (the first one in Table~\ref{cds}) is 0.9~mag, which for a normal interstellar medium extinction law corresponds to 1.6~mag of extinction in the V~band. The reddening associated with this amount of extinction would be E(B-V)~$\sim$ 0.5~mag, which would predict a \ion{Na}{1} column density of log~$N_{\rm NaI}$ =~13.2~cm$^{-2}$ \citep{bohlin78}, much higher than the \ion{Na}{1} column density that we actually detect.
Our interpretation of these facts is that the particulate matter in the KH~15D disk has grown large enough to become both optically grey and dynamically separated from the gas layer. This requires particle sizes of a millimeter or so, consistent with the minimum grain size derived by \citet{herbst08} on the basis of a completely different argument.
It would be interesting to know what the gas-to-dust ratio in the settled particulate layer is. Unfortunately, converting from \ion{Na}{1} column density to total gas column density is difficult for two reasons. First, the ionization fraction of sodium atoms is unknown, and second, it is likely that much of the sodium is depleted onto the solids. If ISM values of the ionization and depletion applied, then there would be significantly more matter in the form of solids than gas. If the gas density is very high, on the other hand, one would expect the solid grains to quickly lose momentum through friction with the gas and spiral into the stars. Unless there were a steady source of new dust, such a structure would disappear on an astronomically brief time scale. It may be, therefore, that the very existence of the particulate disk indicates a fairly low gas to dust ratio.
\section{Summary and Conclusions} \label{summary}
\ion{Na}{1} lines in the spectrum of the binary WTTS KH~15D have been analyzed in detail. We find an excess absorption component that may be attributed to the gas component of a circumbinary disk. The derived column density is log~$N_{\rm NaI}$ $\sim$~12.5~cm$^{-2}$ with no significant variation associated with the position of Star~A relative to the optically occulting edge. Subtracting the likely contribution of the ISM leaves log~$N_{\rm NaI} \sim$~12.3~cm$^{-2}$.
There is no detectable change in the gas column density across the ``knife edge" formed by the edge of the grain disk, indicating that the gas and solids have very different scale heights, with the solids being highly settled.
If a standard ISM ratio of $N_{\rm NaI}$ to A$_{\rm V}$ is applied along these lines of sight, there would be a much lower mass of gas than there is mass of solids in this disk. However, the conversion from \ion{Na}{1} column density to total gas column density is complicated by the fact that much of the Na in this disk gas might be either ionized or depleted onto the solids.
An additional complication is detectable emission in the \ion{Na}{1} feature that follows the radial velocity of Star~A. While this might be due to the slight mismatch in the spectral type of the comparison spectrum, it more likely arises from the active, accreting magnetosphere known to be associated with this (quasi) WTTS.
Our data support a picture of this circumbinary disk as being composed of a very thin particulate grain layer composed of sand (millimeter) sized objects or larger, which have settled within whatever remaining gas may be present. This phase of disk evolution has been hypothesized to exist as a prelude to the formation of planetesimals through gravitational fragmentation \citep{safronov69, goldreich73} but has not been extensively observed, and is expected to be short-lived if much gas were still present in such a disk.
Both components of KH~15D are now fully eclipsed by the circumbinary disk for the forseeable future \citep{chiang04}, so our chance to directly measure absorption in this particular disk has passed. Fortunately, other young, eclipsing, edge-on disk systems such as WL4 \citep{plavchan08} and YLW~16A (Plavchan et~al., in prep.) are being discovered in photometric monitoring surveys, and future spectroscopic studies may allow better understanding of the timescale of dust and gas settling in transition disks.
\acknowledgements{We thank Coryn Bailer-Jones for some of the VLT/UVES data, and Suzan Edwards for the spectrum of LkCa~7. S.~M.~L.\ would like to thank Roy Kilgard and Peter Plavchan for very helpful advice and discussions regarding this paper. W.~H.\ acknowledges support by NASA through the Origins of the Solar Systems grant NNX08AK35G. C.~M.~J.-K.\ acknowledges partial support by NASA through Origins of the Solar Systems grant NNX08AH86G. J.~N.~W.\ gratefully acknowledges the support of the MIT Class of 1942 Career Development Professorship. J.~A.~J.\ acknowledges support from NSF grant AST-0702821.
Some of the data presented herein were obtained at the W.~M.\ Keck Observatory, which is operated as a scientific partnership among the California Institute of Technology, the University of California and the National Aeronautics and Space Administration. The Observatory was made possible by the generous financial support of the W.~M.\ Keck Foundation. The authors wish to recognize and acknowledge the very significant cultural role and reverence that the summit of Mauna Kea has always had within the indigenous Hawaiian community. We are most fortunate to have the opportunity to conduct observations from this mountain. The Hobby-Eberly Telescope (HET) is a joint project of the University of Texas at Austin, the Pennsylvania State University, Stanford University, Ludwig-Maximilians-Universit{\"a}t M{\"u}nchen, and Georg-August-Universit{\"a}t G{\"o}ttingen. The HET is named in honor of its principal benefactors, William P.\ Hobby and Robert E.\ Eberly. This paper is based in part on observations collected at the European Organisation for Astronomical Research in the Southern Hemisphere, Chile (Program 074.C-0604A).}
{\it Facilities:} \facility{HET (HRS)}, \facility{Keck:I (HIRES)}, \facility{VLT:Kueyen (UVES)}
\bibliographystyle{apj}
|
\section{Introduction}
The development of Scanning Tunnelling Microscopy (STM)
recently enabled to study the case of localised magnetic excitations
at surfaces~\cite{Heinrich,Hirjibehedin06,Hirjibehedin07,Tsukahara,XiChen,Iacovita,Fu}.
Low-temperature STM experimental studies
revealed that, in certain cases, a local spin could be associated with
individual adsorbates at surfaces. Under the variation of the STM bias,
the tip-adsorbate junction exhibits conductance steps at well-defined
energies in the few meV energy range. These steps are attributed to the
excitation of the local spin of the adsorbate and the step energy position
yields the corresponding excitation energy, leading to magnetic IETS
(Inelastic Electron Tunnelling Spectroscopy). Revealing the existence
of nano-magnets at the atomic (molecular) level can have fascinating
consequences for the miniaturisation of electronic devices. In addition,
the possibility to determine the energy spectrum of the local spin on
the adsorbate as a function of an applied magnetic B field provides
an efficient way to quantitatively characterize the local spin and
its interaction with the underlying substrate. This makes magnetic
IETS an invaluable tool for magnetic studies of nano-objects at
surfaces. In this respect, one can mention several spectacular results
obtained in this way: evidence of spin coupling between neighbouring
atomic adsorbates (anti-ferromagnetic coupling along adsorbed Mn
chains)~\cite{Hirjibehedin06}, large magnetic anisotropy of atomic adsorbates (Mn and
Fe adsorbates on CuN)~\cite{Hirjibehedin07}, change of magnetic structure of a molecular
adsorbate on various substrates (Fe-Phthalocyanine on Cu and CuO)~\cite{Tsukahara},
evidence of super-exchange interactions~\cite{XiChen}, interaction of molecular
adsorbates with magnetic substrates~\cite{Iacovita}, {charging
of adsorbed magnetic nano-objects~\cite{Fu}}. Magnetic transitions could be
easily evidenced in these systems due to the presence of a coating on
the metal substrate that efficiently decouples the adsorbate from the
continua of metallic states. When magnetic atoms are directly adsorbed
on a metal, magnetic transitions could be observed but much broadened
by the interaction with the substrate~\cite{Balashov}.
Besides detailed insight into
the structure of a magnetic adsorbate on a surface, IETS experiments
bring also information on the dynamics of the excitation of a local
spin by tunnelling electrons. Indeed, when considering a conductance
spectrum as a function of the junction bias, the position of the
conductance steps yields the excited state energies and the height
of the conductance steps relative to each other and relative to the
conductance at zero bias yields the relative magnitude of the various
possible magnetic excitations. This feature is quite important for
possible future applications, since it determines how easily a local
adsorbate spin can be flipped at will by tunnelling electrons or can
be quenched by collisions with substrate electrons~\cite{FDN}. It appears that
magnetic transitions are highly
probable~\cite{Hirjibehedin06,Hirjibehedin07,Tsukahara,XiChen},
much more probable than
other inelastic processes like e.g. vibrational excitation by tunnelling
electrons~\cite{Stipe,Ho,Komeda}. As an extreme example, in the case of Co-Phthalocyanine
(Co-Pc)~\cite{XiChen}, the inelastic contribution to the current is found to be three
times larger that the elastic current, whereas for vibrational excitation,
the inelastic tunnelling contribution reaches at most a few per cent~\cite{Stipe,Ho,Komeda}.
\begin{figure}
\includegraphics[width=0.45\textwidth]{./fig0.eps}
\caption{
Scheme of an
M-Phthalocyanine molecule. In the present
study M is taken as a Fe or a Co atom, here represented in red.
The free molecule
has a $D_{4h}$ symmetry
which is generally reduced upon adsorption on a substrate.
The
N-atoms are depicted in blue, C atoms in grey
and H atoms in cyan.
}
\label{figure0}
\end{figure}
A few tools have been used to describe the spectroscopy and dynamics of a
local spin on an adsorbate. First, the energy of a local spin interacting
with its environment and possibly with an applied magnetic field, B,
has been modelled efficiently with the following magnetic Hamiltonian~\cite{Yosida}:
\begin{equation}
H = g \mu_{B} \vec{B} \cdot \vec{S} + D S_{z}^{2} + E (S_{x}^{2}-S_{y}^{2}),
\label{hamiltonien}
\end{equation}
where $\vec{S}$ is the local spin of the
adsorbate, $g$ the Land\'e factor and $\mu_B$ the Bohr magneton. $\vec{B}$ is an
applied magnetic field. $D$ and $E$ are two energy constants describing
the interaction of $\vec{S}$ with the substrate, i.e. the magnetic anisotropy of
the system. The Cartesian axis xyz are chosen according to the magnetic
symmetries of the system. Hamiltonian~(\ref{hamiltonien})
describes how the adsorbate
spin, $\vec{S}$, is oriented in space due to its interaction with the substrate. The
observed magnetic excitation energies of adsorbates were very efficiently
modelled by the above Hamiltonian~\cite{Hirjibehedin06,Hirjibehedin07,Tsukahara}.
As for the strength of the
magnetic excitations, it was first analyzed in a phenomenological way
by Hirjebehidin {\em et al}~\cite{Hirjibehedin07}.
They showed that the relative heights of the
inelastic conductance steps were very close to the relative magnitude
of the squared matrix elements of the operator between the initial and
final states of the transition. Nothing could be said about the relative
magnitude of the elastic and inelastic conductance. This finding was
later supported by several theoretical
studies~\cite{Fransson,Fernandez,Persson} introducing in
the description of electron tunnelling a coupling term proportional
to the local spin. In first order perturbation theory, the inelastic
current then appears proportional to the squared matrix elements of the
coupling term between initial and final states, similarly to what was
noticed in Ref.~[\onlinecite{Hirjibehedin07}]. Recently, a completely
different approach was introduced to
treat magnetic excitations by tunnelling electrons~\cite{Lorente}.
It is based on the
large difference of time scale between electron tunnelling and magnetic
anisotropy of adsorbates: electron tunnelling is fast and the magnetic
anisotropy can be considered as non-active during tunnelling. One can
then describe tunnelling without the magnetic anisotropy taken into
account and simply switch the latter at the beginning and at the end of
tunnelling. In addition, DFT calculations on the two studied systems (Fe
and Mn adsorbates on CuN) showed that only one coupling scheme between
adsorbate and tunnelling electron spins is significantly contributing
to tunnelling. The magnetic excitation appears as the result of a spin
decoupling/recoupling process. This approach is non-perturbative and
can then handle the large excitation probabilities encountered in these
systems as well as make predictions for the elastic/inelastic relative
contributions to tunnelling. In addition to bringing a qualitative view
into the magnetic excitation process, it was shown to accurately account
for the strength of the magnetic excitations, in particular relative to
the elastic channel, in the case of Mn and Fe atomic adsorbates on CuN~\cite{Lorente}.
In the present work, we show how our method can be used in the
case of Fe-Pc and Co-Pc molecules (Fig.~\ref{figure0}) adsorbed
on partly insulating
substrates. As one of the main results, this approach is shown to
precisely account for the overwhelming dominance of the inelastic
conductance over the elastic conductance observed experimentally
in these systems~\cite{Tsukahara,XiChen}.
\begin{figure}
\includegraphics[angle=0,width=0.45\textwidth]{./fig1.eps}
\caption{
Projected density of states (PDOS) on the Fe $d$-atomic orbitals. For all
the curves shown here, the positive (black) curves corresponds to the
majority spin, and the negative (red) to the minority spin.
The $d$ orbitals are classified according to the cartesian
axes that contain the N--Cu--N axis of the molecule $(x,y,z)$, or
with respect to the surface directions: $x'$ for the $[1-10]$ and $y'$
for the $[001]$ directions. The $z$ axis is the same for both
reference frames.
}
\label{figure1}
\end{figure}
\section{Method}
\label{2}
We consider a local spin, $\vec{S}$, localized on an adsorbate on a surface.
It is coupled to its environment by the Hamiltonian~(\ref{hamiltonien}). Diagonalizing
the Hamiltonian~(\ref{hamiltonien}) yields the various spin states of the system, their
energies, $E_n$, and the associated wave-functions, $\phi_n$.
Hamiltonian~(\ref{hamiltonien})
is more easily written in the basis of $|S,M\rangle$ states, the eigenstates of $\vec{S}^2$
and $S_z$, so that the anisotropy states can be written as:
\begin{equation}
\left|\phi_{n}\right\rangle = \sum_{M} C_{n,M} \left|S,M\right\rangle
\label{phi_n}
\end{equation}
In the present study, we used the $D$, $E$ and $g$ parameters in
Hamiltonian~(\ref{hamiltonien}) as determined in the analysis of the
experimental energy spectrum and from these, without any further parameter
adjustment, we derive the strength of the magnetic transitions induced
by tunnelling electrons. The various energy terms in Hamiltonian~(\ref{hamiltonien})
are in the few meV range, so that one does not expect the corresponding interaction
to play a role during the electron tunnelling process and one can treat
this problem in the sudden approximation: treat the tunnelling without
the magnetic anisotropy and then introduce it as a frame transformation
at the beginning and at the end of tunnelling.
Electron tunnelling from the STM tip through the adsorbate
and into the substrate (in the absence of magnetic anisotropy) can be
represented by a scattering $T$ matrix, noted $T_{Tip\rightarrow Sub}$
(and an equivalent one for the reverse tunnelling).
It corresponds to the scattering of the
tunnelling electron by the adsorbate and it depends on the electron
energy. In the absence of magnetic anisotropy, it depends a priori on the
spin coupling between the tunnelling electron and the adsorbate, via the
exchange interaction. Thus, in the absence of significant spin-orbit interactions, it can be written in a diagonal form if we
consider $\vec{S}_T$, the total spin of the system (electron + adsorbate). Defining
$\left|S_T,M_T\right\rangle$ as
the eigenfunctions of $\vec{S}_T^2$ and $S_{T,z}$ (if $S$ is the adsorbate spin,
then $S_T = S \pm \frac{1}{2}$), we can write
formally the scattering $T_{Tip\rightarrow Sub}$ matrix
(in the absence of magnetic anisotropy) as:
\begin{equation}
T_{Tip\rightarrow Sub}= \sum_{S_T,M_T} \left|S_T,M_T\right\rangle
T_{Tip\rightarrow Sub}^{S_T} \left\langle S_T,M_T\right|.
\label{Tmatrix}
\end{equation}
$T_{Tip\rightarrow Sub}^{S_T}$ is a number,
function of the electron energy. However, since, below,
we consider only the limited energy range spanned by the magnetic
excitations (up to 10 meV), the transmission probabilities $| T_{Tip\rightarrow Sub}^{S_T}|^2$ can be
considered as constant in the present study. In the sudden approximation,
the tunnelling amplitude (in the presence of magnetic anisotropy) is
written as the matrix element of the $T_{Tip\rightarrow Sub}$ amplitude between the initial and
final states of the tunnelling process. These states are written as
$|\frac{1}{2}, m; \phi_n \rangle $ where the
first part concerns the tunnelling electron (the electron spin is 1/2 and
$m$ is the projection of the tunnelling electron spin on the quantization
axis) and the second part concerns the local spin of the adsorbate.
One then obtains the
amplitude, $AMP_{m,n\rightarrow m',n'}$, for a tunnelling electron induced transition from $\phi_n$ to
$\phi_{n'}$, while the tunnelling electron spin projection changes from $m$ to $m'$ as:
\begin{eqnarray}
&_{m,n\rightarrow m',n'} =
\sum_{S_T} \; T_{Tip\rightarrow Sub}^{S_T} & \nonumber \\
&\times
\sum_{M_T} \langle \frac{1}{2}, m'; \phi_n' | S_T,M_T \rangle
\langle S_T,M_T | \frac{1}{2}, m; \phi_n \rangle&
\label{amp}
\end{eqnarray}
One can see that there is a
superposition (interference) of tunnelling amplitudes
through the different $S_T$ introduced by the magnetic anisotropy.
Equation~(\ref{amp})
yields the transition amplitudes in the general case, if the spin
of the tunnelling electron is registered in both the initial and final
states. Implicitly, it has been assumed above that the tunnelling electron
quantization axis is the z-axis of the adsorbate magnetic anisotropy;
situations with different quantization axis for the adsorbate and the tunnelling electron can be easily handled with an expression similar to Eq.~ (\ref{amp}).
We can now define the probability, $P$, for transitions from $\phi_n$ to $\phi_{n'}$
induced by unpolarized tunnelling electrons by summing incoherently over
the distinguishable channels:
\begin{eqnarray}
P_{n\rightarrow n'} &=&
\frac{1}{2} \sum_{m,m'} | AMP_{m,n\rightarrow m',n'}|^2 \nonumber \\
&=& \frac{1}{2} \sum_{m,m'} | \sum_{S_T} T_{Tip\rightarrow Sub}^{S_T} \\
&\times&
\sum_{M_T} \langle \frac{1}{2}, m'; \phi_n' | S_T,M_T \rangle
\langle S_T,M_T | \frac{1}{2}, m; \phi_n \rangle
|^2. \nonumber
\label{prob}
\end{eqnarray}
We can be a little more explicit by writing
the total spin states,$|S_T,M_T\rangle$,
as expansions over uncoupled spin states:
\begin{equation}
|S_T,M_T\rangle = \sum_m CG_{S_T,M_T,m} |S,M=M_T-m\rangle |\frac{1}{2}, m\rangle
\label{SM}
\end{equation}
where $ |S,M\rangle$ states
correspond to the adsorbate spin states and $|\frac{1}{2}, m\rangle$
to the tunnelling electron spin. Again, the adsorbate
and electron spins are quantized on the same axis. The CG are Clebsch-Gordan coefficients.
Combining (\ref{SM}) and (\ref{phi_n}) we
can express the total spin states as an expansion over uncoupled products
of adsorbate magnetic anisotropy states and tunnelling electron spin:
\begin{equation}
\left|j\right\rangle =\left|S_{T},M_{T}\right\rangle =
\sum_{n,m} A_{j,n,m} \left|\phi_{n}\right\rangle \left|1/2,m\right\rangle
\label{Ajnm}
\end{equation}
Equation~(\ref{Ajnm}) links the $ \left|j\right\rangle =\left|S_{T},M_{T}\right\rangle$, states appropriate to describe tunnelling without
magnetic anisotropy to the channel states of the complete tunnelling
process. We can then rewrite the transition probability (\ref{prob}) as:
\begin{equation}
P_{n\rightarrow n'}
= \frac{1}{2} \sum_{m,m'} | \sum_{S_T} T_{Tip\rightarrow Sub}^{S_T}
\sum_{M_T} A_{j,n,m} A^*_{j,n',m'}|^2.
\label{P2}
\end{equation}
In the case where the $T_{Tip\rightarrow Sub}^{S_T}$
tunnelling amplitude is dominated by a single symmetry, $S_T$
(like it was found in the case of Mn and Fe adsorbates on CuN~ \cite{Lorente},
and like it is shown below to be the case in the systems studied here),
the probability (\ref{P2}) further simplifies into:
\begin{equation}
P_{n\rightarrow n'}
= \frac{1}{2} | T_{Tip\rightarrow Sub}^{S_T}|^2 \sum_{m,m'} |
\sum_{M_T} A_{j,n,m} A^*_{j,n',m'}|^2.
\label{P3}
\end{equation}
This result, used in Ref.~[\onlinecite{Lorente}], is very simple,
the electronic part of
the tunnelling (the $T_{Tip\rightarrow Sub}^{S_T}$ amplitude) is factored out and the probabilities
for the different channels are simply proportional to spin-coupling
coefficients corresponding either to the magnetic anisotropy or to
the coupling between electron and adsorbate spins (the coefficients
are products of the diagonalization expansion coefficients in Eq.~ (\ref{phi_n})
and Clebsch-Gordan coefficients).
From the transition probabilities we can write the conductance $dI/dV$ as
a function of the STM bias, $V$, as:
\begin{equation}
\frac{dI}{dV}=C_0\frac{\sum_n \Theta(V-EX_n) \sum_{m,m'}
|\sum_j A_{j,1,n} A^*_{j,n,m'}|^2}{
\sum_n \sum_{m,m'}|\sum_j A_{j,1,n} A^*_{j,n,m'}|^2}.
\label{conductance}
\end{equation}
Expression~(\ref{conductance}) corresponds to the conductance for the
system being initially in the ground state $n=1$. The sum over $n$ extends over all the
$| \phi_n \rangle $ states, including the ground state,
so that the above conductance
takes
{all contributions, elastic and inelastic,}
into account. $EX_n$ is the excitation energy
of the magnetic level $n$, corresponding
to the eigenvalue difference of the final, $| \phi_n \rangle $,
and initial $| \phi_1 \rangle $ states. The Heavyside function, $\Theta$,
takes care of the opening of the inelastic channels at zero temperature.
$ C_0$ is the total
conductance corresponding to the transmission amplitude
$T_{Tip \rightarrow Sub}^{S_T}$.
It is then
a magnetism-independent conductance. Since we only consider a limited $V$
range, defined by the magnetic excitation energies, $C_0$ can be considered
as constant in the relevant $V$-range.
$C_0$ is equal to the conductance of the
system for biases larger than all the inelastic thresholds.
Expression~(\ref{conductance})
corresponds to the case where only one $S_T$ value actually contributes
to tunnelling so that the sum over $j$ is restricted to the corresponding
$M_T$ values. If the two $S_T$ symmetries contribute to tunnelling, a more
general expression derived from Eq.~ (\ref{P2}) has to be used.
The above form (\ref{P3}) for the transition probabilities accounts for the
strength of the magnetic transitions. In fact, the transition probability
is not proportional to a coupling term as in perturbation approaches,
it appears as a simple
{sharing}
of the global transmission of the
tip-substrate junction, $|T_{Tip \rightarrow Sub}^{S_T}|^2$,
among the various spin states, $|\phi_n\rangle$.
Qualitatively,
in the magnetic excitation process, one can say that the incident
electron arrives in a given spin state, it couples with the adsorbate
spin to form a state of the total spin, $S_T$; tunnelling through the
junction occurs independently in the different total spins; at the end
of the electron-adsorbate collision, the total spin splits back into its
adsorbate and electron components, populating all the possible adsorbate
spin states. Then, the expression for the sharing process (\ref{P3}) simply
expresses angular-momentum conservation.
This kind of excitation process has been invoked in various situations
where an exchange of angular momentum is involved: resonant rotational
excitation in electron collisions on free and
adsorbed molecules~ \cite{Abram,Teillet2000},
spin-forbidden transitions in electron-molecule
collisions~\cite{Teillet1987} or in
atom-surface scattering~\cite{Bahrim}.
In all cases, it leads to efficient excitation
processes. One can stress that this description of magnetic transitions is
at variance with that of the vibrational excitation induced
by collisional electrons ~\cite{Gadzuk,Djamo} or by tunnelling
electrons~\cite{Lorente2000,Lorente2004,Paulsson}.
However, even in the vibrational excitation
case, a process similar to the one discussed here for magnetic transitions does
exist. It is associated with the composition of the incident
electron linear momentum with that of the target atom. Conservation of
momentum leads to recoil of the target induced by electron scattering;
however, due to the large ratio
between electron and nuclei masses,
the corresponding excitation is very weak and leads to very small
excitation probabilities. Vibrational excitation has then to involve
other processes such as e.g. resonant scattering, where the increase of
the collision time allows weak interactions to be
efficient in various situations of electron-molecule collisions~\cite{Schulz,Birtwistle,Gadzuk,Djamo,Manip, Domcke,Monturet}. The
above discussion can be summarized by stressing that the angular momenta
(orbital or spin) of electrons and atoms or molecules can be of the same
order of magnitude due to quantization (a few units in many cases) so
that exchange between them is easy and has visible effects; in contrast,
because of the large electron-nuclei mass ratio,
the electron linear
momentum is usually much smaller than that of heavy particles limiting
the efficiency of momentum exchange processes. {As a consequence,} angular
and spin degrees of freedom appear similar and the present treatment of
spin transitions is very similar to the treatment of rotational excitation
used in Ref.~[\onlinecite{Teillet2000}]
to account for the experimentally observed efficiency of
tunnelling electron in inducing molecular adsorbate rotations~\cite{Stipe2}.
\begin{figure}
\includegraphics[width=0.45\textwidth]{./fig2.eps}
\caption{
Amplitude of the Kohn-Sham orbital of the Fe-Pc/Cu(110)(2$\times$1)-O +
STM tip system at 0.2 eV above the Fermi level (positive in light grey
and negative in light pink). The STM tip in the upper part is made of a
Cu(110) surface with an extra protruding atom (in gold). The adsorbed
Fe-Pc is lying flat on the substrate (N atoms in blue, C atoms in grey
and H atoms in cyan). The substrate is in the lower part of the figure
(Cu atoms in gold and O in red). The orbital is concentrated around
the Fe atom and exhibits a strong $d_{z^2}$ character perturbed by the
interaction with the substrate.
}
\label{figure2}
\end{figure}
\section{Magnetic excitations in supported F\lowercase{e}-phthalocyanine molecules}
\label{3}
Recently Tsukahara {\em et al}~\cite{Tsukahara}
performed a detailed magnetic IETS study of
single Fe-Phthalocyanine (Fe-Pc) molecules adsorbed on a Cu(110)(2$\times$1)-O
surface. The molecule lay flat on the surface and two adsorption
geometries were found, labelled $\alpha$ and $\beta$
differing by the relative
orientation of the molecule on the substrate. Clear magnetic transitions
were observed by scanning the STM tip bias and were attributed to a local
spin $S=1$ interacting with the environment and with an applied magnetic
field, $B$. The magnetic transitions were only observed when the tip was
placed above the Fe atom and were attributed to a local spin of the Fe
atom. The energies of the magnetic levels of the system as a function
of the applied field were very precisely accounted for in
Ref.~[\onlinecite{Tsukahara}] using
Hamiltonian~(\ref{hamiltonien}).
The parameters of the two, $\alpha$ and $\beta$ adsorption
geometries are different: $D = -3.8$ meV, $E = 1.0$ meV and $g = 2.3$ for
$\alpha$-Fe-Pc and $D = -6.9$ meV, $E = 2.1$ meV and $g = 2.4$ for
$\beta$-Fe-Pc,
i.e. the same kind of structure but with a very different zero-field
splitting of the magnetic states. We used this modelling of the magnetic
structure to compute the strength of the magnetic excitations as
induced
by tunnelling electrons, making use of the formalism described in
section~\ref{2}.
\subsection{Electronic structure of the Fe-phthalocyanine molecule}
\label{molecule}
A detailed description of the electronic structure of the Fe-Pc can be
found in various references (see e.g. [\onlinecite{Dale,XLu,Liao}]
and earlier references there in).
The $s$ outer electrons of the Fe atom are transferred to the
surrounding Pc ring leaving a central Fe$^{2+}$ ion with a $3d^6$ electronic
configuration. In the free Fe-Pc, see Fig.~\ref{figure0},
the Fe atom is surrounded by the
Pc ring of D$_{4h}$ symmetry, so that the Fe $d$ manifold splits into $b_{2g}$
($d_{xy}$), $e_{g}$ ($d_{zx},d_{yz}$), $a_{1g}$ ($d_{z^2}$) and
$b_{1g}$ ($d_{x^2-y^2}$) orbitals (the $z$-axis is normal to the Pc
plane
{and the $x$ and $y$-axis are parallel to the Fe-N interatomic axis}
). In the free Fe-Pc, the $b_{2g}$
($d_{xy}$) orbital is fully occupied, the $b_{1g}$ ($d_{x^2-y^2}$)
corresponds to the highest eigenenergy because of the
large overlap with the N-atom orbitals,
and four electrons occupy the $e_{g}$ ($d_{zx},d_{yz}$) and
$a_{1g}$ ($d_{z^2}$)
orbitals. Various configurations have been proposed, with small
energy differences between them (see a discussion in Ref.~[\onlinecite{Liao}]). The structure
of the other M-Pc (M= metal) is similar~\cite{Liao},
a change of M along the Fe,
Co, Ni, Cu and Zn sequence corresponding to the further filling of the
split $d$ manifold.
When the Fe-Pc is adsorbed on a Cu(110)(2$\times$1)-O surface,
the D$_{4h}$ symmetry is broken. If we assume the presence of the surface to
only induce a perturbation on the structure of the free Fe-Pc, then the
$g$ subscript of the orbitals disappears and the $e_g$ orbital splits (the
$ d_{zx}$
and $d_{yz}$ are not degenerate anymore). We will
use the orbital notation of the free
Fe-Pc when
discussing the electronic
structure of adsorbed Fe-Pc, although the molecules
are distorted by the interaction with the surface.
\begin{figure}
\includegraphics[angle=0,width=0.45\textwidth]{./fig3.eps}
\caption{
In panel (a), the transmission function for the system shown
in Fig.~\ref{figure2}
as a function of the tunnelling electron energy (referred to the Fermi energy $E_F$) for a vanishing STM bias.
Panel (b) shows the PDOS on the Fe $d_{z^2}$ atomic orbital.
Panel (c) shows the projection onto the $a_{1g}$ molecular orbital of the free distorted Fe-Pc
that corresponds to the $d_{z^2}$ Fe orbital.
Positive (black) curves correspond to the majority spin,
and the negative (red) ones to the minority spin.
}
\label{figure3}
\end{figure}
\subsection{Density functional study of Fe-Pc adsorbed on a Cu(110)(2$\times$1)-O
surface}
\label{surface}
The ground state electronic structure configuration and the global
transmission, $T_{Tip \rightarrow Sub}^{S_T}$
for the tunnelling from an STM tip { to the substrate} passing through
the Fe-Pc molecule were obtained by density functional theory (DFT)
simulations, performed with
the {\sc Siesta} and {\sc Transiesta} codes~\cite{Siesta,Transiesta}.
We have used the generalized gradient approximation~\cite{PBE}
for the exchange-correlation functional. We only studied the $\beta$ adsorption geometry
of the Fe-Pc on Cu(110)(2$\times$1)-O, which has a higher symmetry than the $\alpha$ geometry.
The electronic structure in the $\alpha$ and $\beta$ geometries are assumed to be equivalent.
Starting with the electronic structure analysis for the adsorbed molecule,
{\sc Siesta} uses atomic orbitals for the basis set, and the projection of the
density of states onto the Fe atomic orbitals
reveals {an open shell structure with two unpaired electrons, i.e.} a $S=1$ configuration
($d_{xy}^2 d_{y'z}^2 d_{z^2}^1 d_{x'z}^1$), as shown
in Fig.~ \ref{figure1}.
Note that the $\beta$ adsorption geometry corresponds to the Fe-N axis rotated by 45$^\circ$
from the symmetry axis of the Cu(110) surface. As a consequence, the splitting of the $e_g$ orbital
by the substrate field involves $x'$ and $y'$-axis rotated by 45$^\circ$ from the $x$ and $y$-axis that
correspond to the splitting of the $d$ manifold by the interaction with the Pc ring. The projections in Fig.~\ref{figure1}
were then performed on the appropriate symmetry orbitals i.e. the $d_{xy}$, $d_{z^2}$, $d_{x^2-y^2}$, $d_{zx'}$ and $d_{y'z}$ orbitals.
More quantitatively,
the computed spin polarization is 1.85 unpaired electrons. The corresponding
spin is
roughly $S=1$, in good agreement with the experiment~\cite{Tsukahara}.
Due to its large overlap with the surface electronic structure,
the $d_{z^2}$ orbital is the one that hybridizes the most with
the substrate.
\begin{figure}
\includegraphics[width=0.45\textwidth]{./fig4.eps}
\caption{
Transmission eigenchannel corresponding to the
largest transmission amplitude for
(a) $E=E_F$ where $E_F$ is the system's Fermi energy and
(b) $E=E_F + 0.2$ eV. Light grey (light pink) color corresponds to the positive (negative)
imaginary part of the eigenchannel amplitude coming from the STM tip.
In gold color, the positive real part. Note that the isosurfaces were chosen different in
(a) and (b) because of the large difference in transmission probability.
The transmission channel exhibits a strong $a_{1g}$ ($d_{z^2}$) character around the Fe atom.
}
\label{figure4}
\end{figure}
Figure~\ref{figure2}
shows the real space plot of the Kohn-Sham orbital at 0.2 eV above the Fermi level for the
molecule+surface+STM tip system. This orbital corresponds to the peak in the PDOS onto the $d_{z^2}$
Fe orbital (see Fig.~\ref{figure1}). The strong $d_{z^2}$ character of the orbital seen in
Fig.~\ref{figure2} confirms that this orbital corresponds to the $a_{1g}$ orbital perturbed by the substrate.
The intuitive notion that, by reaching further out along the $z$-axis, this orbital would
contribute more to the junction transmission is
supported by comparing the transmission function
(panel (a)
of Fig.~\ref{figure3})
with the
PDOS onto the $d_{z^2}$ Fe atomic orbital (b), and PDOS onto the molecular
orbital $a_{1g}$ of the Fe-Pc (c).
Note that,
for this projection, we used the $a_{1g}$ orbital of the free Fe-Pc molecule,
with the distorted atomic configuration of the adsorbed molecule.
From this, it is clear that, at the Fermi level, the transmission is
essentially made of the tail of the $a_{1g}$ resonance due to the $d_{z^2}$ Fe orbital,
and that only the minority spin contributes to the tunneling tranmission.
In the transmission calculations, the tip-molecule distance was set at a small distance, 5~\AA, with a large enough transmission probability to visualize the electron transmission eigenchannel easily.
This conclusion is further confirmed
by computing the transmission eigenchannels. For that, we have used
the {\sc Inelastica} code~ \cite{Inelastica}. The $d_{z^2}$ orbital is the largest contributor to S-matrix
eigenchannel {dominating} the
transmission of electrons from the STM tip (here
represented by an atom on a Cu(110) semi-infinite electrode)
placed above the
molecule.
As can be seen in Fig.~\ref{figure4}, the character of the
{ dominant transmission} eigenchannel is the same at the Fermi energy, $E_F$, Fig.~\ref{figure4}(a)
and at $E=E_F + 0.2$ eV, Fig.~\ref{figure4} (b),
the peak of the transmission resonance. These plots nicely correspond to the
density plot of the Kohn-Sham eigenstate, Fig.~\ref{figure2}.
\begin{figure}
\includegraphics[width=0.45\textwidth]{./fig5.eps}
\caption{
Computed conductance for the (a) $\alpha$ and (b) $\beta$ configuration
of an adsorbed Fe-Pc molecule on a Cu(110)(2$\times$1)-O
surface.
The conductance has been normalized to 1 at large bias
and the curves for the various $B$ fields have been offset
by 0.5 for representation purposes.
The B field is oriented vertical to the sample with
magnitude 0, 3, 6, 9 and 11 T as indicated in the graph.
\label{figure5}}
\end{figure}
\subsection{Orbital picture of the electron tunnelling process}
From the above DFT study, we can conclude that the Fe-Pc on a
Cu(110)(2$\times$1)-O surface,
is associated to the triplet electronic configuration
$d_{xy}^2 d_{y'z}^2 d_{z^2}^1 d_{x'z}^1$
consistently with experiment~ \cite{Tsukahara}. The
interaction between the $d$ manifold and the surroundings is seen
to split completely the $d$ manifold, leading to five non-degenerate
orbitals; the orbital angular momentum is then completely quenched,
in the absence of significant spin-orbit couplings~\cite{Yosida}
and this justifies
the discussion, used in Ref.~\cite{Tsukahara} and here,
of the magnetic anisotropy in
terms of the spin angular momentum orientation. One can also notice that
the Fe spin is also partially quenched from its free atom value ($S =
2$), again due to the interaction with the surroundings that induces a
large upward energy shift of the $d_{x^2-y^2}$ orbital.
The present DFT study also
shows that when an STM tip is placed above the Fe atom, tunnelling
dominantly involves the $d_{z^2}$ ($a_{1g}$) orbital. So when an electron is sent
from the tip on the Fe, it involves the {$d_{xy}^2 d_{y'z}^2 d_{z^2}^2 d_{x'z}^1$}
{transient} configuration, i.e. the total
spin of the electron-adsorbate scattering intermediate is $S_T = 1/2$.
Similarly, if a hole
is sent from the tip to the Fe
atom, tunnelling involves the {$d_{xy}^2 d_{y'z}^2 d_{z^2}^0 d_{x'z}^1$} {transient}
configuration, i.e. the total spin of the
electron-adsorbate scattering intermediate is again $S_T = 1/2$. Thus, of
the two possible symmetries for the electron-adsorbate scattering ($S_T =
1/2$ or $3/2$),
$S_T = 1/2$ is the prevailing
symmetry in the tunnelling process.
{
In addition, transmission through the $d_{z^2}$ orbital will dominate
in a constant-current STM image and
generate a bright spot at the Fe centre}. This is consistent with the
observation in Ref.~[\onlinecite{Tsukahara}]
of the Fe-Pc molecule as a bright spot at the centre
of a clover leaf. The bright spot corresponds to
the $d_{z^2}$ Fe orbital and the clover leaf is given by
the contribution from other orbitals localized on the Pc ring (see in~[\onlinecite{XLu}]
a discussion of the link between bright M atoms in M-Pc STM images and their
$d$ orbitals). Furthermore, the magnetic transitions in this system were
also found~\cite{Tsukahara} to be localized in the same region
about the Fe atom, {which we attribute to the $d_{z^2}$ orbital.}
\subsection{Magnetic excitation processes}
The inelastic conductance of the Fe-Pc has been computed using
expression~(\ref{conductance}) for the $\alpha$ and $\beta$
adsorption geometries, using the spin
parameters determined in the DFT study. The conductance is shown
in Fig.~\ref{figure5}(a) for $\alpha$ configuration
and Fig.~\ref{figure5} (b) for the $\beta$ configuration, as a function of the STM bias
for various values of the applied
magnetic field, $B$, along the $z$-axis. The conductance has been {normalized to 1} at large bias
and the curves for the various $B$ fields have been {offset
by 0.5}. In the calculation, a Gaussian broadening of 0.25 meV
was introduced to mimic various broadening effects. The conductance
spectra resemble very much those measured by Tsukahara
{\em et al}~\cite{Tsukahara}, with
well-marked steps at the magnetic excitation thresholds (two inelastic
thresholds for this $S=1$ system because anisotropy splits
the triplet state in three states~\cite{Tsukahara})
and very significant contributions
from the inelastic currents. Indeed, the conductance is dominated by
inelastic tunnelling at large bias!
The differences between the $\alpha$ and $\beta$
adsorption geometries are also well reproduced. One can also stress
that, in the present approach, the magnetic excitations do not modify the
maximum value of the global conductance $C_0$,
which is then independent of changes
in the magnetic structure and in particular independent of the applied $B$
field. This appears clearly in Figs.~\ref{figure5} (a) and (b)
as well as in Fig.2a and 2b of Ref.~[\onlinecite{Tsukahara}],
confirming our view of the magnetic excitation process as a sharing of
the global conductance over the various possible magnetic channels.
The relative magnitude of the elastic and inelastic channels in this
system are further illustrated in Fig.~\ref{figure7}
which presents the magnitude
of the inelastic conductance steps for
$C_0=1$.
The two inelastic step heights are $\alpha_1$ and $\alpha_2$ ($\alpha_1$
for the lowest threshold) for the $\alpha$ geometry (and similarly for
the $\beta$ one).
The elastic conductance is then equal to the global conductance
minus the inelastic ones, (1 - $\alpha_1$ - $\alpha_2$)
with this definition.
The present theoretical results as functions of
the applied $B$ field are compared with the data of
Tsukahara {\em et al}~\cite{Tsukahara}. As
a first remark, inelastic tunnelling contributes significantly to the
tunnelling current: typically for $B = 0$ T, the inelastic current is
equal to twice the elastic current. Second, we can see that the present
theoretical results reproduce very well the {relative magnitude of the three contributions to the
conductance (elastic and two inelastic).}
In particular, the variation with $B$ is well accounted for.
The variation of the strength of the magnetic excitation with $B$ is in fact
reflecting the change of the magnetic structure of the adsorbate. For $B
= 0$, the structure of the conductance curve
is due to the anisotropy imposed by the substrate to
the Fe-Pc molecule; although the level positions are different in the
$\alpha$ and $\beta$
adsorption geometries, the eigenstates of Hamiltonian (\ref{hamiltonien})
at $B = 0$ are
the same in both geometries (see Tsukahara {\em et al}~\cite{Tsukahara})
and consequently,
the excitation probabilities are the same, as seen in Fig.~\ref{figure7}.
The effect of a finite $B$ field is to decouple the Fe-Pc spin
from the substrate and
to tend to a Zeeman limit at large $B$. This decoupling is easier for the
$\alpha$-geometry, because of a weaker transversal
anisotropy $E$ (given above). In the limit of very
large $B$ (not fully reached here), the $\phi_n$ states reduce to
$|S,M\rangle$ states with
the ground state corresponding to $M = -1$. The only allowed transition is
to the ($M = 0$) excited state, so that only one inelastic step remains
in the conductance spectrum. Its height is given by
the modulus square of a Clebsch-Gordan
coefficient and in this limit, the inelastic current is equal to one
third of the total current.
\begin{figure}
\includegraphics[width=0.45\textwidth]{./fig6.eps}
\caption{
Relative inelastic step heights in the conductance
as a function of the magnetic field B, for (a) $\alpha$ and
(b) $\beta$ configurations. $\alpha_1$ and $\alpha_2$
refer to the first and second excitation steps for the $\alpha$ configuration
respectively. Analogously, $\beta_1$ and $\beta_2$ refer
to the first and second excitation steps for the $\beta$ configuration.
The experimental data points are represented
with black squares for the first excitation and
as red circles for the second one and are taken from the supplemental
material of Ref.~[\onlinecite{Tsukahara}].
\label{figure7}}
\end{figure}
\section{Magnetic transitions in supported C\lowercase{o}-Phthalocyanine molecules}
\label{CoPc}
{
Superposed layers of Co-Pc molecules on Pb(111) islands were studied
experimentally by magnetic IETS and revealed the existence of
superexchange interactions~\cite{XiChen}}. Magnetic
excitation of the Co-Pc molecules
were observed but only for several adsorbed layers
of Co-Pc molecules. No excitation was observed in the single layer case,
suggesting that
either the spin of the Co-Pc molecular layer
lying directly on the Pb surface
was quenched by the interaction with the substrate or the possible
magnetic
excitations were very short lived~\cite{XiChen}.
The Co-Pc spin was found to be $S = 1/2$,
and when several layers are stacked one on top of each other,
with a stacking angle of 60$^\circ$, the spins of the Co-Pc molecule
in the second and
outer layers couple together in a way well-described by a Heisenberg
Hamiltonian, with an anti-ferromagnetic exchange coupling, $J$, of the
order of 18 meV~\cite{XiChen}.
Very strong magnetic excitations were reported in the
multi-layer case~\cite{XiChen}. The
present formalism can be used to predict the magnitude of the magnetic
excitation in such systems. If we neglect the magnetic anisotropy of
adsorbed Co-Pc, the only ingredients in our approach are the adsorbate
spin $S$, the intermediate total spin $S_T$, and the anti-ferromagnetic
interaction $J$ for the stacked molecules. Both $S$ and $J$ are
known from experiment~\cite{XiChen}.
We did not perform a detailed DFT study as for
the Fe-Pc case.
We assumed that we can {extrapolate the electronic structure from Fe-Pc to Co-Pc~\cite{Liao}},
in both cases the Pc molecules being partly decoupled from the underlying metal.
As discussed in Ref.~\onlinecite{Liao},
an M-Pc series is formed by the various incomplete
$d$-shell metals and the Co structure corresponds to adding a $d$ electron
to the Fe case. With the open shell structure of Fe-Pc outlined above
{($d_{xy}^2 d_{y'z}^2 d_{z^2}^1 d_{x'z}^1$)},
this leads to a doublet configuration of the Co-Pc electronic structure,
independently of the orbital
in which the electron is added. This is perfectly consistent with the
experimental observation. Then, we can assume that tunnelling involves
minority spin electrons because
the molecular electronic structure at the Fermi level is given by
the last partially-occupied $d$ orbital also in this case. Hence, the
intermediate spin state is $S_T = 0$, both for electrons and holes.
In the case of a single active Co-Pc molecule (two adsorbed layers),
an applied magnetic field is needed to perform a magnetic IETS
experiment, because of the molecular spin $S = 1/2$,
see Ref.~[\onlinecite{XiChen}]. For a finite $B$
field, a Zeeman splitting corresponding to a Land\'e factor 1.88 has
been observed~ \cite{XiChen}.
Assuming a simple Zeeman structure, i.e. no magnetic
anisotropy induced by the substrate, the above formalism, at finite $B$,
predicts a single inelastic peak with an inelastic current equal to the
elastic current (note that this result is independent of the value of the
Land\'e factor). This is consistent with the observations~\cite{XiChen},
which reported
an increase of the current around 110-120 \% at the inelastic threshold
(see Fig.~2c in Ref.~[\onlinecite{XiChen}]).
In the case of three molecular layers, two stacked molecules in the
upper two layers interact via an anti-ferromagnetic coupling, so that the
ground state of the two-molecule system is a spin 0 state while the spin
1 states are excited states (we neglect a possible anisotropy induced by
the substrate). Magnetic excitations of the system by tunnelling electrons
for a vanishing $B$ have been observed experimentally as a sharp step in the
conductance~\cite{XiChen}.
The above formalism predicts at $B = 0$, in the absence of any
anisotropy induced by the substrate, a single conductance step at finite
bias (the transition from $S = 0$ to $S = 1$), the inelastic current being
three times larger than the elastic one. An accurate quantitative comparison
with experiment of the step height is difficult because of the non-flat
behaviour of the global conductance in this system (see Ref.~[\onlinecite{XiChen}]); nevertheless, our
prediction of a 300 \% increase of the conductance at the inelastic threshold is
in good agreement with the experimental data. For finite $B$, the $S=1$
excited states split into a Zeeman structure and the present formalism
predicts three steps in the conductance, each of them with a height
equal to the elastic conductance. Again, this Zeeman limit corresponds
to what is observed experimentally for large $B$ (see e.g. Fig.~3b in
Ref.~[\onlinecite{XiChen}]).
As the main feature of this system, one can stress the
extremely large contribution of the inelastic current compared to that
of the elastic current in the total current:
the present theory predicts a total inelastic
current three times larger than the elastic one, in good agreement
with what is observed experimentally. One can also emphasize that in the
Co-Pc case, we only discuss the Zeeman structure limit, so that the only
ingredients in our approach are the value of the local spin ($S = 1/2$)
and of the spin of the electron+adsorbate system ($S_T = 0$) and all the
predictions of relative magnitudes of elastic and inelastic currents
are directly obtained from Clebsch-Gordan coefficients.
\section{Summary and concluding remarks}
The orientation of the magnetic moment of individual adsorbates on a
surface leads to a magnetic structure with excitation energies in the
few meV range. Recent experimental developments in the low-Temperature
Inelastic
Electron Tunnelling Spectroscopy allowed the direct observation of
transitions among these magnetic states in the case of magnetic atoms
adsorbed on a metal with a decoupling coating in between. The present
paper reports a theoretical study of magnetic excitations induced
by tunnelling electrons in metal-phthalocyanine molecules adsorbed on
surfaces (Fe-Pc on Cu(110)(2$\times$1)-O and Co-Pc stacked on Co-Pc on Pb)
that have been the subject of recent experimental studies~\cite{Tsukahara,XiChen}. It is
based on a theoretical framework recently introduced~\cite{Lorente}
to treat tunnelling
electron-induced magnetic transitions and {on a DFT calculation to characterize the electronic structure of the Fe-Pc molecule on Cu(110)(2$\times$1)-O system.}
Our approach determines the strength of the magnetic transitions
induced by tunnelling electrons when the STM tip is placed on top
of the magnetic atom. The input ingredients in our calculations are
the magnetic Hamiltonian describing the interaction of the adsorbate
magnetic moment with its environment and the total spin of the tunnelling
electron-adsorbate system. The magnetic Hamiltonian has been taken from
its parameterisation using the experimental results on the energy spectrum
of magnetic levels~\cite{Tsukahara,XiChen}.
In the Fe-Pc on Cu(110)(2$\times$1)-O case, a DFT study
determined the electronic structure of the adsorbed Fe-Pc molecule
together with
the symmetry and spin structure of the tunnelling electron. An excellent
account of the experimental findings was obtained; in particular, the
extremely large magnetic excitation probabilities (inelastic contribution
dominating over the elastic one in the tunnelling current) were confirmed.
The Fe atom in the adsorbed Fe-Pc has electronic and magnetic structures quite different from those of a free Fe atom and of Fe adsorbed on CuN studied earlier~\cite{Hirjibehedin07,Lorente}. The interaction of the $d$-manifold with the Pc ring and with the CuO substrate results in a full splitting of the manifold and in a spin state different from the atomic case. Then a Fe atom inside an adsorbed Fe-Pc molecule~\cite{Tsukahara} appears very differently in a magnetic IETS experiment compared to the case of an adsorbed Fe atom~\cite{Hirjibehedin07}.
The magnetic transitions appear to be much more probable than other
inelastic processes studied earlier, such as vibrational excitation
of the adsorbate~\cite{Stipe,Ho,Komeda}.
In the most spectacular case (Co-Pc, see above and
Ref.~[\onlinecite{XiChen}]),
the inelastic contribution to the tunnelling current is three
times larger than the elastic one. The present approach explains this
striking difference. Since electron tunnelling occurs on a short time
scale compared to magnetic anisotropy, one can treat electron tunnelling
independently of the magnetic anisotropy in the sudden approximation. The
magnetic transitions then appear as the result of a change of coupling
scheme for the adsorbate spin: coupling to the adsorbate environment in
the initial and final states and coupling to the tunnelling electron
spin via exchange interactions for the tunnelling process. The strength
of the magnetic transitions is then determined by spin coupling
coefficients (such as e.g. Clebsch Gordan coefficients) and the present
approach reduces to computing how a magnetism-independent tunnelling
current is shared among the various magnetic states, i.e. how a total
magnetism-independent current is shared between elastic and inelastic
parts. The importance of a given magnetic transition is then linked to
the weight of the initial and final states in the magnetism-independent
collision intermediate and it can thus be very large. In particular, it
does not depend on the strength of an interaction coupling initial and
final states during electron tunnelling. Our approach is, thus, perfectly
well-adapted to treat situations like the present ones, where tunnelling
appears to be dominated by inelastic effects.
The strength of the magnetic transitions thus appears to be the direct
consequence of the spin coupling scheme of the system. The variation
of the magnetic transitions with an applied magnetic field, $B$, follows
the variation of the adsorbate magnetic structure with $B$, basically the
switch from a magnetic anisotropy induced by the adsorbate environment
to a Zeeman structure, i.e. the decoupling of the adsorbate spin from
its environment by the $B$-field action (see e.g. Fig.~\ref{figure5}).
In this way,
the present study of the magnetic transitions strength as a function of $B$
further strengthens the knowledge of the adsorbate magnetic structure as
it can be derived from the analysis of the experimental energy spectrum
of the magnetic states. Basically, the analyses of the strength of the
magnetic transitions and of the energy spectrum are probing the same
properties of the system.
\begin{acknowledgments}
We are grateful to Prof. Noriaki Takagi and Prof. Maki Kawai for
extended discussions on the experimental data. We thank Prof. Magnus
Paulsson for his help in some of the calculations.
Financial support from the spanish MICINN through grant FIS2009-12721-C04-01
is gratefully ackowledged. N.L. thanks Universit\'e Paris-Sud for an
Invited Professorship and for its hospitality. F.D.N. would like to thank the
Centro de Supercomputaci\'on de Galicia (CESGA) for providing computational resources.
\end{acknowledgments}
|
\section{Introduction}
Let $Z$ be a locally compact metric space, $\mathcal L$ a
$p$-dimensional
lamination on $Z$. {\em We assume throughout that $\mathcal L$ is
minimal}. Let $h: \mathbb R^p\times X\to Z$ be a lamination chart,
i.\ e.\ a homeomorphism onto an open subset $h(\mathbb R^p\times X)$
such that the plaque $h(\mathbb R^p\times\{x\})$
lies on a leaf of $\mathcal L$ for any $x\in X$. We identify $X$ with
the image $h(\{0\}\times X)$ and call it a {\em cross section} of
$\mathcal L$. With the metric induced from $Z$, $X$ is also locally
compact. Notice that any leaf of $\mathcal L$ intersects $X$.
Given a leafwise curve joining two points $x$ and
$y$
on $X$, a holonomy map along $c$ is defined as usual
to be a local homeomorphism $\gamma$ from an
open neighbourhood ${\rm Dom}(\gamma)$ of $x$ onto an open neighbourhood
${\rm Range}(\gamma)$ of $y$. We say that $\mathcal L$ is {\em transversely
equicontinuous} w.\ r.\ t.\ a cross section $X$ if the
family of all the corresponding holonomy maps is equicontinuous.
\begin{theorem} \label{t1}
Let $\mathcal L$ be a minimal lamination
on a locally compact metric space $Z$, transversely
equicontinuous
w.\ r.\ t.\ a cross section $X$. Then there is a Radon measure
on $X$ which is left invariant by any holonomy map. If further $Z$ is compact, then the invariant measure is
unique up to a scaling.
\end{theorem}
The existence of invariant measure was already shown by R.
Sackesteder in \cite{S} for a pseudogroup acting on a compact metric
space. But the compactness condition for a cross section
is too strong
to obtain a corresponding result for laminations or foliations
(even on compact spaces or manifolds).
In section 2, we include a slightly general theorem applicable to laminations;
the proof closely follows an argument in Lemme 4.4 in \cite{DKN},
which is meant for codimension one
foliations.
In section 3 we show the uniqueness for a compact lamination. The
argument here which is adapted for pseudo*groups as defined in section
2
is rather messy, but the
original idea is quite simple, which the reader can find in section 4.
In section 4, we deal with an equicontinuous group action on a compact
metric
space, together with a random walk on a group. We show that the
corresponding harmonic probability measure on the space is unique.
The uniqueness of harmonic measures for tangentially sufficiently
smooth foliations and laminations (\cite{C},\cite{G}) remains an open question.
\section{The existence}
Let $Y$ be a Hausdorff space. By a local homeomorphism, we
mean a homeomorphism $\gamma$ from an open subset ${\rm Dom}(\gamma)$
of $Y$ onto an open subset ${\rm Range}(\gamma)$.
A set $\Gamma$ of local homeomophisms of $Y$ is called a {\em pseudo*group},
if it satisfies the following conditions.
\\
(1) If $\gamma\in\Gamma$ and $U$ is an open subset of ${\rm Dom}(\gamma)$,
then the restriction $\gamma\vert_U$ is in $\Gamma$.
\\
(2) The identity $id_X$ belongs to $\Gamma$.
\\
(3) If $\gamma,\gamma'\in\Gamma$ and ${\rm Dom}(\gamma')={\rm Range}(\gamma)$,
then the composite $\gamma'\circ\gamma$ is in $\Gamma$.
\\
(3) If $\gamma\in\Gamma$, then $\gamma^{-1}\in\Gamma$.
This differs from the usual definition of pseudogroups in
that it does not assume the axiom for taking the union.
Thus for example the set of all the holonomy maps w.\ r.\ t.\
a cross section given in section 1 forms a pseudo*group,
while the pseudogroup they generate might be bigger.
There are two reasons for introducing the concepts of pseudo*groups:
one is that in Theorem \ref{t1}, assuming the equicontinuity
for the pseudogroup generated by the holonomy maps may be
stronger than what we have tacitly in mind: the other is that
some part of the argument in section 3 cannot be put into
the framework of the usual pseudogroups.
\bigskip
Let $X$ be a locally compact metric space and $\Gamma$ a pseudo*group
of local homeomorphsims of $X$. We assume that
the action is minimal, i.\ e.\ the $\Gamma$-orbit of any point is dense
in $X$, and that the action is equicontinuous, i.\ e.\ for any
$\epsilon>0$, there is $\delta(\epsilon)>0$ such that if
$\gamma\in\Gamma$, $x,x'\in{\rm Dom}(\gamma)$ and $d(x,x')<
\delta(\epsilon)$,
then we have $d(\gamma x,\gamma x')\leq \epsilon$.
Denote by $C_c(X)$ the space of real valued continuous functions $\zeta$
whose support ${\rm supp}\zeta$ is compact.
A Radon measure $\mu$ on $X$
is called $\Gamma$-invariant if whenever $\zeta\in C_c(X)$ and
$\gamma\in\Gamma$ satisfy ${\rm supp}\zeta\subset{\rm Dom}(\gamma)$, we have
$\mu(\zeta\circ\gamma^{-1})=\mu(\zeta)$.
In fact if $\mu$ is $\Gamma$-invariant, we get a bit more, e.\ g.\
for any bounded continuous function $\zeta: X\to{\mathbb R}$ which
vanishes outside ${\rm Dom}(\gamma)$, we have
$\mu(\zeta\circ\gamma^{-1})=\mu(\zeta)$,
as the dominated convergence theorem shows. In this case the both
hand sides might be $\infty$. This will be used
in section 3.
Let $X_0$ be
a relatively compact open subset of $X$, and denote by $\Gamma_0$
the restriction of $\Gamma$ to $X_0$ i.\ e.\
$$
\Gamma_0=\{\gamma\in\Gamma\mid {\rm Dom}(\gamma)\cup {\rm Range}(\gamma)\subset X_0\}.$$
The purpose of this
section is to show the following theorem.
\begin{theorem} \label{t2}
There exists a finite $\Gamma_0$-invariant Radon measure $\mu$ on $X_0$.
\end{theorem}
The minimality assumption shows then the existence of $\Gamma$-invariant
measure on $X$ and the proof of the existence part of Theorem \ref{t1}
will be complete.
\newcommand{C_c(X)_{\geq0}}{C_c(X)_{\geq0}}
\newcommand{C_c(X_0)_{\geq0}}{C_c(X_0)_{\geq0}}
\newcommand{C_c(X)_{>0}}{C_c(X)_{>0}}
\newcommand{C_c(X_0)_{>0}}{C_c(X_0)_{>0}}
\bigskip
Let us define
$$
C_c(X)_{\geq0}=\{\zeta\in C_c(X)\mid \zeta\geq 0\}\ \ \ \rm{and}
$$
$$
C_c(X)_{>0}=\{\zeta\in C_c(X)_{\geq0}\mid \zeta(x)>0, \ \ \exists x\in X\}.
$$
For any $\psi\in C_c(X)$ and $\gamma\in\Gamma$, extend the function
$\psi\circ\gamma^{-1}$ to the whole $X$ so as to vanish outside
${\rm Range}(\gamma)$ and still denote it by $\psi\circ\gamma^{-1}$. It may
no longer be continuous.
For any $\zeta\inC_c(X)_{\geq0}$ and $\psi\inC_c(X)_{>0}$, define $(\zeta:\psi)$ by
$$
(\zeta:\psi)=\inf\{\sum_{i=1}^nc_i\mid\zeta\leq\sum_{i=1}^nc_i
\psi\circ\gamma_i^{-1},\ c_i>0,\ \gamma_i\in\Gamma,\ n\in{\mathbb N} \}.
$$
Notice that the minimality of $\Gamma$ implies that
$(\zeta:\psi)<\infty$
and $(\zeta:\psi)=0$ if and only if $\zeta=0$.
Fix once and for all a function $\chi\inC_c(X)_{>0}$ such that $\chi=1$ on $X_0$,
and define a map $L_\psi:C_c(X)_{\geq0}\to{\mathbb R}$ by
$$
L_\psi(\zeta)=(\zeta:\psi)/(\chi:\psi).
$$
It is routine to show the following properties of $L_\psi$.
\begin{eqnarray}
L_\psi(c\zeta)=cL_\psi(\zeta),\ \forall c\geq0,\\
L_\psi(\zeta_1+\zeta_2)\leq L_\psi(\zeta_1)+L_\psi(\zeta_2),\\
\zeta_1\leq\zeta_2\Rightarrow L_\psi(\zeta_1)\leq L_\psi(\zeta_2),\\
{\rm supp}\zeta\subset{\rm Dom}(\gamma)\Rightarrow L_\psi(\zeta\circ\gamma^{-1})
=L_\psi(\zeta),\\
L_\psi(\zeta)\geq 1/(\chi:\zeta).
\end{eqnarray}
\begin{lemma} \label{l1}
If $\eta>0$ and $\xi, \xi'\inC_c(X)_{\geq0}$ satisfies $\xi+\xi'=\chi$, then
there is $\delta>0$ such that if $\psi\inC_c(X)_{>0}$, ${\rm diam}({\rm supp}\psi)<\delta$
and
$\zeta\inC_c(X_0)_{\geq0}$ we have
$$
L_\psi(\xi\zeta)+L_\psi(\xi'\zeta)\leq (1+2\eta)L_\psi(\zeta).
$$
\end{lemma}
\noindent
{\bf Proof.} Given $\eta>0$, there is $\epsilon>0$ such that if
$x, x'\in X_0$ and $d(x,x')\leq \epsilon$, then
$\abs{\xi(x)-\xi(x')}\leq \eta$.
Also this implies $\abs{\xi'(x)-\xi'(x')}\leq \eta$.
Choose $\delta=\delta(\epsilon)$. Let
$\psi$ be as in the lemma and assume
\begin{equation} \label{e1}
\zeta\leq \sum_{i}c_i\psi\circ\gamma_i^{-1}.
\end{equation}
Notice that if we restrict $\gamma_i$ in (\ref{e1}) to
${\rm Dom}(\gamma_i)
\cap{\rm supp}\psi\cap\gamma_i^{-1}({\rm supp}\zeta)$, still the inequality
(\ref{e1})
holds. Hence if we choose $x_i$ from ${\rm Range}(\gamma_i)\subset
{\rm supp}(\zeta)\subset X_0$,
then for any $x\in{\rm Range}(\gamma_i)$, we have
$$
\abs{\xi(x)-\xi(x_i)}\leq\eta\ \ {\rm and} \ \ \abs{\xi'(x)-\xi'(x')}\leq \eta.
$$
Moreover the following inequality
$$
\xi(x)\psi\circ\gamma_i^{-1}(x)\leq(\xi(x_i)+\eta)\psi\circ\gamma_i^{-1}(x)
$$
holds for any $x\in X$, since if $x\not\in{\rm Range}(\gamma_i)$ the both
hand
sides are 0.
Then we have
\begin{eqnarray*}
\zeta(x)\xi(x)\leq\sum_ic_i\xi(x)\psi\circ\gamma_i^{-1}(x)\\
\leq\sum_ic_i(\xi(x_i)+\eta)\psi\circ\gamma_i^{-1}(x).
\end{eqnarray*}
This shows
$$
(\zeta\xi:\psi)\leq\sum_ic_i(\xi(x_i)+\eta).
$$
We have a similar inequality for $\xi'$. Since
$x_i\in X_0$ and thus $\xi(x_i)+\xi'(x_i)=1$, we have
$$
(\zeta\xi:\psi)+(\zeta\xi':\psi)\leq(2\eta+1)\sum_ic_i.
$$
The lemma follows from this.
\hfill
q.\ e.\ d.
\bigskip
Continuing the proof of Theorem \ref{t1}, let us extend the operator
$L_\psi:C_c(X_0)_{\geq0}\to\mathbb R$ to $C_c(X_0)$ by just putting
$$
L_\psi(\zeta)=L_\psi(\zeta_+)-L_\psi(\zeta_-),
$$
where $\zeta_+$ (resp.\ $\zeta_-$) is the positive (resp.\ negative) part
of $\zeta$.
Then we have:
\begin{equation}
\abs{L_\psi(\zeta)}\leq \Vert \zeta\Vert_\infty, \ \ \forall \zeta\inC_c(X_0)_{\geq0}.
\end{equation}
In fact if $\zeta\geq 0$, then $\zeta\leq\Vert\zeta\Vert_\infty\chi$,
and thus $L_\psi(\zeta)\leq\Vert\zeta\Vert_\infty$, the general case
following easily from this.
Let us identify $L_\psi$ with the following point of a compact
Hausdorff space:
$$
L_\psi=\{L_\psi(\zeta)\}_\zeta\in\prod_{\zeta\in C_c(X_0)}
[-\Vert\zeta\Vert_\infty,\Vert\zeta\Vert_\infty].
$$
Let $\{\psi_n\}$ be a sequence in $C_c(X)_{>0}$
such that ${\rm diam}({\rm supp}\psi_n)\to0$.
Choose an operator
$L\in\bigcap_{m}{\rm Cl}\{L_{\psi_n}\mid n\geq m\}$.
This means that for any finite number of elements $\zeta_\nu\in C_c(X_0)$
and any $\epsilon>0$, there is
a sequence $n_i\to\infty$ such that $\abs{L(\zeta_\nu)-L_{\psi_{n_i}}
(\zeta_\nu)}<\epsilon$. Now we have the following properties of
the map $L:C_c(X_0)\to\mathbb R$.
\begin{eqnarray}
L(c\zeta)=cL(\zeta),\ \ \forall c\in{\mathbb R},\\
L(\zeta_1+\zeta_2)\leq L_(\zeta_1)+L(\zeta_2),
\ \ \forall \zeta_1,\ \zeta_2\geq 0, \label{e0}\\
\zeta_1\leq\zeta_2\Rightarrow L(\zeta_1)\leq L(\zeta_2),\\
{\rm supp}\zeta\subset{\rm Dom}(\gamma),\ \gamma\in\Gamma_0\Rightarrow L(\zeta\circ\gamma^{-1})
=L(\zeta),\label{e2}\\
\zeta\in C_c(X_0)_{>0} \Rightarrow L(\zeta)\geq 1/(\chi:\zeta), \label{e3}\\
\abs{L_\psi(\zeta)}\leq \Vert\zeta\Vert_\infty \label{e5}.
\end{eqnarray}
Moreover by
Lemma \ref{l1} and (\ref{e0}), we have
\begin{lemma} \label{l2}
If $\zeta\in C_c(X_0)_{\geq0}$ and $\xi,\ \xi'\in C_c(X)_{\geq0}$ satisfy $\xi+\xi'=\chi$,
then
$$
L(\xi\zeta)+L(\xi'\zeta)=L(\zeta).
$$
\end{lemma}
From this one can derive the linearity of $L$. First of all notice that
\begin{equation} \label{e4}
\zeta,\zeta'\inC_c(X_0)_{\geq0}\Rightarrow \abs{L(\zeta)-L(\zeta')}\leq\Vert\zeta
-\zeta'\Vert_\infty.
\end{equation}
In fact we have
\begin{eqnarray*}
L(\zeta')=L(\zeta+\zeta'-\zeta)\leq L(\zeta+(\zeta'-\zeta)_+)\leq
L(\zeta)+L((\zeta'-\zeta)_+)\\
\leq L(\zeta)+\Vert (\zeta'-\zeta)_+\Vert_\infty\leq L(\zeta)+\Vert\zeta'-\zeta
\Vert_\infty.
\end{eqnarray*}
Continuing the proof of the linearity, notice that it suffices to show
it
only for those functions $\zeta_1, \zeta_2\in C_c(X_0)_{\geq0}$.
Choose $\epsilon>0$ small and let
$$\xi_j=(\zeta_j+\epsilon\chi)/(\zeta_1+\zeta_2+2\epsilon)
$$ for $j=1,2$.
Then we have $\xi_1+\xi_2=\chi$. Now
$$
\xi_1(\zeta_1+\zeta_2)-\zeta_1=\epsilon(\zeta_2-\zeta_1)/(\zeta_1
+\zeta_2+\epsilon).
$$
Therefore by (\ref{e4}), we have
$$
\abs{L(\xi_1(\zeta_1+\zeta_2))-L(\zeta_1)}\leq\epsilon.
$$
On the other hand by Lemma \ref{l2},
$$
L(\xi_1(\zeta_1+\zeta_2))+L(\xi_2(\zeta_1+\zeta_2)))
=L(\zeta_1+\zeta_2).
$$
Since $\epsilon$ is arbitrary, we have obtained
$$
L(\zeta_1)+L(\zeta_2)=L(\zeta_1)+L(\zeta_2),$$
as is requied.
\bigskip
Now $L$, being a positive operator, corresponds to a Radon measure
$\mu$.
By (\ref{e3}), the measure $\mu$ is nontrivial, and since
(\ref{e5})
implies
$$
\inf\{L(\zeta)\mid \zeta\in C_c(X_0)_{\geq0},\ \Vert\zeta\Vert_\infty\leq1\}\leq
1,$$ the measure $\mu$ satisfies $\mu(X_0)\leq 1$.
Finally (\ref{e2}) means the $\Gamma_0$-invariance of $\mu$.
\section{The uniqueness}
In this section $\Gamma$ is again an equicontinuous
and minimal pseudo*group of local homeomorphisms
of a locally compact metric space $X$. The modulus of equicontinuity
is also denoted by $\epsilon\to\delta(\epsilon)$. Denote by
$B_r(x)$ the open $r$-ball in $X$ centered at $x\in X$.
We make the following additional assumption on the pseudo*group $\Gamma$.
\begin{assumption} \label{a1}
There is a relatively compact open subset $X_0$ of $X$ and $a>0$
such that if $\gamma\in\Gamma$, $x\in X_0$, $x\in{\rm Dom}(\gamma)\subset B_a(x)$
and $\gamma x\in X_0$, then there is $\hat\gamma\in\Gamma$ such that
${\rm Dom}(\hat\gamma)=B_a(x)$ and $\hat\gamma\vert_{{\rm Dom}(\gamma)}=\gamma$.
\end{assumption}
The purpose of this section is to show the following theorem.
\begin{theorem} \label{t3}
Let $\Gamma$ be an equicontinuous and minimal pseudo*group on $X$
satisfying
Assumption \ref{a1}. Then the $\Gamma$-invariant Radon measure on
$X$ is unique up to a scaling.
\end{theorem}
First of all let us show that the holonomy pseudo*group $\Gamma$ on a cross
section $X$ of a minimal
lamination on a compact space $Z$, equicontinuous w.\ r.\ t.\ $X$
satisfies Assumption \ref{a1}.
Choose any relatively compact open subset $X_0$ of $X$.
On one hand by compactness of $Z$ there is $L>0$ such that the germ of any element
of the restriction $\Gamma_0$ to $X_0$
is a finite composite of the holonomy maps along leaf curves
of length $\leq$ $L$ that join two points in $X_0$.
On the other hand there is $a'>0$ such that each leaf
curve of length $\leq L$
starting at $x\in X_0$ and ending at a point in $ X_0$
admits a holonomy map defined on the ball $B_{a'}(x)$.
An easy induction shows that Assumption \ref{a1} is satisfied
for $a=\delta(a')$.
\bigskip
Let us embark upon the proof of Theorem \ref{t3}.
Choosing $a$ even smaller, one may assume that there is
a nonempty open subset $X_1$ of $X_0$ such that the $a$-neighbourhood
$B_a(x)$ of any point $x$ of $X_1$ is contained in $X_0$ and that if
$\gamma\in\Gamma$ and $x'\in X_0$ satisfies
${\rm Dom}(\gamma)=B_a(x')$ and $\gamma x'\in X_1$,
then the image ${\rm Range}(\gamma)=\gamma(B_a(x'))$ is contained in $X_0$.
Choose $b>0$ so that $b\leq \delta(a/3)$, and assume
there is $x_0\in X_1$ such that $C={\rm Cl}(B)\subset X_1$,
where $B=B_b(x_0)$.
Let $M$ be the space of continuous maps from $C$ to $X_0$,
with the supremum distance $d_\infty$. Define
$$
\Gamma_C=\{\gamma\vert_C \mid \gamma\in \Gamma, \ C\subset{\rm Dom}(\gamma),\
\gamma C\subset X_0\}$$
and let $G$ be the closure of $\Gamma_C$ in $M$.
\begin{lemma} \label{l3}
(1) $G$ is a locally compact metric space.
\\
(2) Any $g\in G$ is a homeomorphism onto a compact subset $gC$ in $X_0$ and
$g$, as well as the inverse map $g^{-1}$, is $\delta(\epsilon)$-continuous.
\end{lemma}
\noindent
{\bf Proof.} All that needs proof is the $\delta(\epsilon)$-continuity
of $g^{-1}$.
Assume $\gamma_n\in\Gamma_C$ converge to $g\in G$ in the $d_\infty$-distance.
If $x,x'\in C$ satisfy $d(x,x')>\epsilon$, then $d(\gamma_nx,
\gamma_nx')\geq\delta(\epsilon)$ by the equicontinuity of the
inverse map $\gamma_n^{-1}$. Thus $d(gx,gx')\geq \delta(\epsilon)$,
as is required.
\hfill q.\ e.\ d.
\bigskip
Recall the notations $B=B_b(x_0)$ and $C={\rm Cl}(B)$.
\newcommand{{\rm Cl}}{{\rm Cl}}
\begin{lemma} \label{l5}
If $g_n\to g$ in $G$, and $y\in gB$, then for any large $n$
we have $y\in g_nB$ and $g_n^{-1}y\to g^{-1}y$.
\end{lemma}
\noindent
{\bf Proof.}
Choose an arbitrary point $x\in B$ and $\epsilon>0$ such that
${\rm Cl}(B_\epsilon(x))\subset B$.
First let us show that for any $\gamma\in \Gamma_C$,
\begin{equation} \label{e11}
B_{\delta(\epsilon)}(\gamma x)\subset \gamma
{\rm Cl}(B_\epsilon(x)).
\end{equation}
In fact, by the choice of the number $b$, we have
$\gamma(B)\subset {\rm Cl}(B_{a/3}(\gamma x_0))$.
That is, $\gamma(B)\subset B_a(\gamma x)$, and thus
$(\gamma\vert_B)^{-1}$ admits an extension $\widehat{\gamma^{-1}}\in\Gamma$
defined on $B_a(\gamma x)$. Choose an arbitrary point
$y\in B_{\delta(\epsilon)}(\gamma x)$. Then by the $\delta(\epsilon)$-continuity
of $\widehat{\gamma ^{-1}}$, the point $x'=\widehat{\gamma^{-1}}y$
lies in ${\rm Cl}(B_{\epsilon}(x))\subset B$. On the other hand
$x'=\gamma^{-1}\gamma x'=\widehat{\gamma^{-1}}\gamma x'$. Since
$\widehat{\gamma^{-1}}$
is injective, we have $y=\gamma x'$. This finishes the proof of
(\ref{e11}).
\bigskip
Next let us show that for any $g\in G$, we have
\begin{equation} \label{e12}
B_{\delta(\epsilon)/2}(gx)\subset g{\rm Cl}(B_\epsilon(x)).
\end{equation}
Again assume $\gamma_n\in\Gamma_C$ converge to $g\in G$.
Since $\gamma_nx\to gx$, we have for any large $n$ that
$B_{\delta(\epsilon)/2}(gx)\subset B_{\delta(\epsilon)}(\gamma_nx)$.
Thus if $y\in B_{\delta(\epsilon)/2}(gx)$, then by (\ref{e11})
$y=\gamma_nx_n$ for some
$x_n\in {\rm Cl}(B_\epsilon(x))$. Passing to
a subsequence, assume that $x_n\to x'\in {\rm Cl}(B_\epsilon(x))$.
Now in the following inequality
$$
d(gx',y)=d(gx',\gamma_nx_n)\leq d(gx',\gamma_nx')
+d(\gamma_nx',\gamma_nx_n),
$$
both terms of the RHS can be arbitrarily small if $n$ is sufficiently
large.
That is, $y=gx'$, showing (\ref{e12}).
\bigskip
To finish the proof of the lemma, assume $g_n\to g \in G$ and $y\in gB$.
By (\ref{e12}), for any sufficiently small $\epsilon>0$
we have
$B_{\delta(\epsilon)/2}(g_ng^{-1}y)\subset g_n{\rm
Cl}(B_\epsilon(g^{-1}y))$.
Since $g_ng^{-1}y\to y$,
we have $y\in g_n{\rm Cl}(B_\epsilon(g^{-1}y))$ for any large $n$
and therefore $g_n^{-1}y\in {\rm Cl}(B_\epsilon(g^{-1}y))$.
Since $\epsilon$ is arbitrarily small, this shows the lemma.
\hfill q.\ e.\ d.
\bigskip
Let $\Gamma_0$ be the restriction of the psudogroup $\Gamma$ to $X_0$.
We shall construct
a pseudo*group $\Gamma_\sharp$ of local homeomorphisms of $G$.
For any $\gamma\in\Gamma_0$, define
\begin{eqnarray*}
& {\rm Dom}(\gamma_\sharp)=\{ g\in G\vert gC \subset {\rm Dom}(\gamma)\}, \\
& {\rm Range}(\gamma_\sharp)=\{ g\in G\vert gC \subset {\rm Range}(\gamma)\}, \\
& \gamma_\sharp g=\gamma\circ g,\ \ \forall g\in {\rm Dom}(\gamma_\sharp.)
\end{eqnarray*}
It may happen that for some $\gamma\in\Gamma_0$,
${\rm Dom}(\gamma)={\rm Range}(\gamma)=\emptyset$. In that case $\gamma_\sharp$
is not defined.
\begin{lemma} \label{l6}
The subsets ${\rm Dom}(\gamma_\sharp)$ and ${\rm Range}(\gamma_\sharp)$ are open in
$G$, and $\gamma_\sharp$ is $\delta(\epsilon)$-continuous w.\ r.\
t.\ the metric $d_\infty$.
\end{lemma}
{\bf Proof.} The easy proof is omitted.
\hfill q.\ e.\ d.
\bigskip
Denote by $\Gamma_\sharp$ the pseudo*group consisting
of all the
elements $\gamma_\sharp$ for $\gamma\in\Gamma_0$ and their
restrictions to open subsets of the domains.
\begin{lemma} \label{l7}
The action of $\Gamma_\sharp$ on $G$ is minimal.
\end{lemma}
{\bf Proof.}
First let us show that for $\gamma_1,\gamma_2\in\Gamma_C$, there is
$\gamma_\sharp\in\Gamma_\sharp$ such that $\gamma_1\in{\rm Dom}(\gamma_\sharp)$
and that $\gamma_\sharp(\gamma_1)=\gamma_2$.
Since $\gamma_1C\subset B_a(\gamma_1x_0)$, there is an element
$\gamma'\in\Gamma$ defined on $B_a(\gamma_1x_0)$ which extends
$\gamma_2\circ\gamma_1^{-1}$. Let $\gamma\in\Gamma_0$ be the restriction
of $\gamma'$ to $\Gamma_0$, i.\ e.\ the restriction such taht
${\rm Dom}(\gamma)=B_a(\gamma_1x_0)\cap
X_0\cap \gamma'^{-1}X_0$. Clearly $\gamma_1C$ is contained in
${\rm Dom}(\gamma)$, showing the claim.
Thus we have shown that $\Gamma_\sharp$-orbit of $id_C$ is nothing but
$\Gamma_C$ and hence dense in $G$. To finish the proof, we shall show
that for any $g\in G$, the $\Gamma_\sharp$-orbit of $g$ visits
an arbitrarily small neighbourhood of any element $\gamma_2\in\Gamma_C$.
Let $\epsilon$ be any small number such that the
$2\epsilon$-neighbourhood
of $\gamma_2C$ is contained in $X_0$. Take $\gamma_1\in\Gamma_C$ such
that
$d_\infty(g,\gamma_1)<\delta(\epsilon)$. Choosing $\epsilon$ and hence
$\delta(\epsilon)$ even smaller, one may very well assume that $gC$
is contained in $B_a(\gamma_1x_0)$.
Then the element
$\gamma\in\Gamma_0$ constructed above (for $\gamma_1$ and $\gamma_2$)
contains $gC$ in its domain, i.\ e.\ $g$ is contained in
${\rm Dom}(\gamma_\sharp)$, and furthermore $d_\infty(\gamma_\sharp g,
\gamma_2)<\epsilon$.
\hfill q.\ e.\ d.
\bigskip
Now by Lemmata \ref{l3}, \ref{l6} and \ref{l7}, one can apply
Theorem \ref{t2} to $(\Gamma_\sharp, G)$ to find a
$\Gamma_\sharp$-invariant
Radon measure $m$ on $G$.
(This is the point where the concept of pseudo*group is useful.
Notice that even if $\Gamma_\sharp$ is equicontinuous,
it does not necessarily imply that
the pseudogroup generated by $\Gamma_\sharp$ is equicontinuous.)
One can assume $m$ is a probability measure
since $G$ is in fact a precompact open subset of a bigger space.
Now let $\mu$ and $\mu'$ be distinct $\Gamma_0$-invariant
probability measures on $X_0$. Then their restrictions to $B$ are
also distinct, by the minimality of the $\Gamma_0$-action.
That is, there is a function $\zeta\in C_c(B)_{}$ such that
$\mu(\zeta)\neq\mu'(\zeta)$. One may assume further that
$\zeta$ is nonnegative valued.
\begin{lemma} \label{l8}
For any $g\in G$, we have
$$
\int_{X_0}\zeta (g^{-1}x)\mu(dx)=\int_{X_0}\zeta(x)\mu(dx).
$$
\end{lemma}
{\bf Proof.} For $g\in\Gamma_C$, this is just the $\Gamma_0$-invariance
of $\mu$. For general $g$, assume $\gamma_n\to g$ for
$\gamma_n\in \Gamma_C$. Then by Lemma \ref{l5}, if $x\in gB$, then
$x\in\gamma_n B$ for any large $n$ and $\gamma_n^{-1}x\to g^{-1}x$.
If $x\not\in gB$, then since $\gamma_n{\rm supp}(\zeta)\to g{\rm supp}(\zeta)$
in the Hausdorff distance, $\zeta(\gamma_n^{-1}x)=0$ for any large $n$,
as well as $\zeta(g^{-1}x)$. In any case for any $x\in X_0$, we have
$\zeta(\gamma_n^{-1}x)\to\zeta(g^{-1}x)$. The lemma follows from
the dominated convergence theorem.
\hfill q.\ e.\ d.
\bigskip
Now recall the space $X_1$. It is an open subset of $X_0$ which
contains $C$ such that
the $a$-neighbourhood
$B_a(x)$ of any point $x$ of $X_1$ is contained in $X_0$
and that if
$\gamma\in\Gamma$ and $x'\in X_0$ satisfies
${\rm Dom}(\gamma)=B_a(x')$ and $\gamma x'\in X_1$,
then the image ${\rm Range}(\gamma)=\gamma(B_a(x'))$ is contained in $X_0$.
\begin{lemma} \label{l9}
The function
$$
Z(x)=\int_G\zeta(g^{-1}x)m(dg)
$$
is constant on $X_1$.
\end{lemma}
{\bf Proof.} Define a function $\zeta_x:G\to {\mathbb R}$ by
$\zeta_x(g)=\zeta(g^{-1}x)$. Lemma \ref{l5} and an additional argument
as above shows that
$\zeta_x$ is a continuous function.
Choose $x,x'\in X_1$ on the same $\Gamma$-orbit.
By the assumption of $X_1$, there is $\gamma\in\Gamma_0$
such that $\gamma x=x'$ and ${\rm Dom}(\gamma)=B_a(x)\subset X_0$ and
${\rm Range}(\gamma)\subset X_0$. Then
we have
$$
\{g\in G\mid\zeta_x(g)>0\}\subset {\rm Dom}(\gamma_\sharp).
$$
In fact if $\zeta_x(g)=\zeta(g^{-1}x)>0$, then $x\in gB$.
On the other hand, diam$(gB)\leq 2a/3$, and thus
$gC\subset B_a(x)={\rm Dom}(\gamma)$, i.\ e.\ $g\in{\rm Dom}(\gamma_\sharp)$.
By the $\Gamma_\sharp$-invariance of the measure $m$, we have
\begin{eqnarray*}
Z(x)=\int_G\zeta_x(g)m(dg)=\int_G\zeta_x(\gamma_\sharp^{-1}(g))m(dg)
=\int_G\zeta_x(\gamma^{-1}\circ g)m(dg)\\
=\int_G\zeta(g^{-1}\gamma x)m(dg)
=\int_g\zeta_{\gamma x}(g)m(dg)=Z(x')
\end{eqnarray*}
That is, the function $Z$ is constant along a $\Gamma$-orbit in
$X_1$. On the other hand it is continuous, since $\zeta\circ g^{-1}$
has the same modulus of continuity. Now the minimality of
$\Gamma_0$-action on $X_1$ shows the lemma.
\hfill q.\ e.\ d.
\bigskip
\begin{lemma} \label{l10}
The function $Z$ is constant on $X_0$.
\end{lemma}
{\bf Proof.} It suffices to show that for
any $x'\in X_0$ and $x\in X_1$ on the same $\Gamma_0$-orbit, we have
$Z(x)=Z(x')$. By the assumption of $X_1$, there exists
an element $\gamma\in\Gamma_0$ such that $\gamma x'= x$ and
${\rm Dom}(\gamma)=B_a(x')\cap X_0$. Then just as before, one can show
$$
\{g\in G\mid\zeta_{x'}(g)>0\}\subset {\rm Dom}(\gamma_\sharp).
$$
Again by the $\Gamma_\sharp$-invariance of $\mu$, we have
$Z(x)=Z(x')$.
\hfill q.\ e.\ d.
\bigskip
Now let us finish the proof of Theorem \ref{t3}. By Lemma \ref{l10},
the function $Z$ is constant on $X_0$, depending only on $\zeta$
and $m$.
We have on one hand
$$
\int_{X_0}\int_G\zeta(g^{-1}x)m(dg)\mu(dx)=\int_{X_0}Z\mu(dx)=Z.
$$
On the other hand by Fubini and by Lemma \ref{l8}
$$
Z=\int_G\int_{X_0}\zeta(g^{-1}x)\mu(dx)m(dg)=\int_G\mu(\zeta)m(dg)=\mu(\zeta).
$$
Since $Z$ does not depend on the choice of $\mu$, we have
$\mu(\zeta)=\mu'(\zeta)$, contrary to the assumption.
\section{The uniqueness of harmonic measures for group actions}
Here the notations of the previous sections are all abandoned.
Let $\alpha:\Gamma\times X\to X$ be an effective (i.\ e.\ faithful) action of
a countable group $\Gamma$ on a
compact metric space $X$, and let $p$ be a probability measure
on $\Gamma$, i.\ e.\ a function $p:\Gamma\to[0,1]$
such that $\sum_{\gamma\in\Gamma}p(\gamma)=1$. We assume
that ${\rm supp}(p)=\{\gamma\in\Gamma\mid p(\gamma)>0\}$
generates $\Gamma$ as a semigroup.
A probability measure $\mu$ on $X$ is called {\em $p$-harmonic}
if $\mu=\alpha_*(p\times\mu)$, that is, for any continuous function
$f$ on $X$, we have
$$
\int_Xf(x)\mu(dx)=\int_X\sum_{\gamma\in\Gamma}p(\gamma)f(\gamma x)\mu(dx).
$$
This section is devoted to the proof of the following theorem.
\begin{theorem} \label{t4}
If the action $\alpha$ is equicontinuous and minimal, then
the $p$-harmonic probability measure $\mu$ on $X$ is unique.
\end{theorem}
{\bf Proof.}
Let $M$ be the space of continuous maps from $X$ to $X$,
endowed with the supremum metric $d_\infty$, and let $G$ be the closure
of $\Gamma$ in $M$.
Then as in section 3, Lemmata \ref{l3} and \ref{l5},
we can show that $G$ is a compact metrizable
topological group, (with the topology induced from the metric
$d_\infty$).
Let $f$ be an arbitrary continuous function on $X$.
Let $m$ be a Haar probability measure on $G$. Define a function
$f_m:X\to {\mathbb R}$
by
$$
f_m(x)=\int_Gf(gx)m(dg).
$$
The function $f_m$ is on one hand continuous since the functions $f\circ g$
have the same modulus of continuity, and on the other hand constant
on $\Gamma$-orbits by the right invariance of $m$. Hence by the
minimality
of the action, $f_m$ is
a constant, which we denote by $c(f, m)$.
Let $\mu$ be a $p$-harmonic probability measure on $X$, and define a function
$f_\mu:G\to {\mathbb R}$ by
$$
f_\mu(g)=\int_Xf(gx)\mu(dx).
$$
Then $f_\mu$ is a continuous function w.\ r.\ t.\ $d_\infty$,
and by the $p$-harmonicity of $\mu$, it satisfies
$$
f_\mu(g)=\sum_{\gamma\in\Gamma}p(\gamma)f_\mu(g\gamma ).
$$
If $f_\mu$ takes the maximal value at $g\in G$, then for
any $\gamma\in{\rm supp}(p)$, the value of $f_\mu$ at $g\gamma $ is also
the maximal. Repeating this arguments one can show that
$f_\mu$ takes the maximal value on the coset $g\Gamma $, since
${\rm supp}(p)$ generates $\Gamma$ as a semigroup.
Because $\Gamma$ is dense in $G$, the function $f_\mu$ must be
a constant, equal to $f_\mu(e)=\mu(f)$.
Now we have
$$
\int_X\int_Gf(gx)m(dg)\mu(dx)=\int_X c(f,m) \mu(dx)=c(f,m).$$
On the other hand by Fubini,
$$
c(f,m)=\int_G\int_Xf(gx)\mu(dx)m(dg)=\int_G\mu(f)m(dg)=\mu(f).$$
Then the value $\mu(f)$, being equal to $c(f,m)$, does not
depened on $\mu$, showing the uniqueness of $\mu$.
\hfill q.\ e.\ d.
\bigskip
|
\section{Introduction}
Let $V$ be a super vector space over a field $k$ of characteristic of $0$. The super group $GL(V)$ of linear automorphisms of $V$ is the
subgroup of the semi-group $\mbox{\rm End\hspace{0.1ex}}(V)$ of endomorphisms with invertible super-determinant. In \cite{Manin1}, Manin introduced the following Koszul
complex $K$ to define the super determinant.
Its $(k,l)$-term is given by \break $K^{k,l} := \Lambda _k\otimes S_l^*$, where $\Lambda_n$ and $S_n$ are the $n$-th homogeneous components of
the exterior and the symmetric tensor algebras on $V$. The differential $d_{k,l}: K^{k,l} \longrightarrow K^{k+1,l+1}$ is given by
$$
d_{k,l} (h\otimes \varphi ) = \sum_i h\wedge x_i \otimes \xi ^i\cdot \varphi .
$$
There is another Koszul complex associated to $V$, denoted by $L$.
This complex was first defined by Priddy as a free resolution of $k$ as a module over the symmetric tensor
algebra of $V$, see \cite{Manin}.
Its $(l,k)$-term is given by $L^{l,k} := S_l\otimes \Lambda_k$ with differential $P_{l,k}: L^{l,k} \longrightarrow L^{l-1,k+1}$ given by
$$ \xymatrix { P_{l,k}: S_l\otimes \Lambda_k \ar@{^(->}[r]&V^{\otimes l}\otimes V^{\otimes k}=V^{\otimes l-1}\otimes V^{\otimes k+1}
\ar[rr]^{\qquad \quad X_{l-1}\otimes Y_{k+1}}&&
S_{l-1}\otimes \Lambda_{k+1} },$$
where $ X_l, Y_k$ are the symmetrizer and anti-symmetrizer operators.
In \cite{Kac3}, Kac proved that any finite dimensional irreducible representation of the Lie super algebra
$\frak{gl}(V)$ is a quotient of the Kac module. He divided irreducible representations of $\frak{gl}(V)$ into two classes,
typical representations and atypical representations.
By using the Kac module, Kac gave an explicit construction of all typical representations of $\frak{gl}(V)$, and
a character formula for all typical representations.
In \cite{Zhang2}, Su and Zhang gave a character formula for all finite-dimensional irreducible representations of $\frak{gl}(V)$. An explicit construction of atypical
representations is however not known.
The aim of this work is to give a combinatorial way to describe
all irreducible representations in case the super-dimension of $V$ is $(3|1)$.
Our observation is that the two Koszul complexes above can be combined into a
double complex which we call the double Koszul complex. We use the differential of this complex to describe all irreducible representations of
$\frak {gl}(V)$ when $V$ has super-dimension $(3|1)$.
The paper is organized as follows. Section $2$ provides some background materials on the general linear super-algebra needed for the rest of the paper.
Section 3 introduces and studies the double Koszul complex. Section 4 uses the properties of the double Koszul complex to construct representations of the
Lie super algebra $\frak {gl}(V)$. Using the character formula of Su and Zhang in \cite{Zhang2}, we prove that the constructed representations furnish all
irreducible representations of $\frak {gl}(V)$.
{\it Acknowledgements:} I would like to thank my teacher, Prof Phung Ho Hai for his constant help and suggestions. I also would like to thank the refree for usefull comments.
\section{Preliminaries}
This section presents some results on the general linear Lie super-algebras for the later use. We shall work with a field $k$ of characteristic $0$.
A super vector space is a \break $\mathbb {Z}_2$-graded vector space $V = V_{\bar 0}\oplus V_{\bar 1} .$
The spaces $ V_{\bar 0}, V_{\bar 1}$
are called even and odd homogeneous components of $V$, their elements are
called homogeneous. We denote the $\mathbb Z_2$-grade (or parity) of a homogeneous element $a$ by $\hat a$.
Assume $\dim V_{\bar 0}=m, \dim V_{\bar 1}=n$ and fix a homogeneous basis of $V$: $x_1,\ldots,x_m \in
V_{\bar 0}$, $ x_{m+1},\ldots,x_{m+n}\in V_{\bar 1}.$ For simplicity we denote the $\mathbb Z_2$-grade of $x_i$ by
$\hat i$. Thus $\hat i=\bar 0$ if $1\leq i\leq m$ and $\hat i=\bar 1$ if $m+1\leq i\leq m+n$.
A $\mathbb Z_2$-graded algebra $A$ is called a super algebra.
Similarly we have the notion of super Lie algebra $L$, where the super anti-commutativity and the super Leibniz rule read:
$$\begin{array}{rcl} [a,b]&=&(-1)^{\hat a\hat b}[b,a],\\[0em]
[a,[b,c]]&=&[[a,b],c]+(-1)^{\hat a\hat b}[b,[a,c]].\end{array}$$
Here we use the convention that $(-1)^{\bar 0}=1$ and $(-1)^{\bar 1}=-1$.
Given a super algebra $A$, the super-commutator on $A$, defined by
$$[a,b]:=ab-(-1)^{\hat a \hat b}ba,$$
makes $A$ into a super Lie algebra, denoted by $A^L$.
\subsection{Super Lie Algebras $\frak g=\frak{gl}(V)$} Consider the algebra $\mathrm{End}(V)$
of linear endomorphisms of $V$.
Fix a homogeneous basis of $V$ as above. Every element of $\mathrm{End}(V)$ is given by a matrix
of the form $ \left({A\ B\atop C\ D}\right)$, where $A,B,C,D$ are block matrices. The matrices of the form
$ \left({A\ 0\atop 0\ D}\right)$ define even maps $V\to V$ (i.e. maps that preserve the $\mathbb Z_2$-grading).
The matrices of the form $ \left({0\ B\atop C\ 0}\right)$ define odd maps (i.e. maps that interchange the $\mathbb Z_2$-grading).
An arbitrary map $V\to V$ is the sum of an even map with an odd map. This
defines a $\mathbb Z_2$-grading on $\mathrm{End}(V)$ and makes $\mathrm{End}(V)$ a super algebra.
The associated super Lie algebra $\mathrm{End}(V)^L$ is denoted by $\frak{gl}(V)$.
\subsection{ Representation of $\frak g= \frak {gl}(V)$}
Let $W$ be a super vector space. A super representation $\rho $ of $\frak g$ in $W$
is an even linear mapping $\rho : \frak g\longrightarrow \frak{gl}(W)$ which preserves the super commutator,
that is a homomorphism of Lie super algebras. A super representation of $\frak g$ is also called a $\frak g$-module. A super representation is said to be irreducible if it has no proper
non-zero sub-representations. In oder to construct all irreducible representations of $\frak g$ we need the
technique of induced representations, which we will now describe.
\subsubsection{Induced representations}
A pair $(\mathcal{U}(\frak g), i)$, where
$ \mathcal{U}(\frak g)$ is an associative $\mathbb{Z}_2$-graded algebra and $i: \frak g \longrightarrow \mathcal{U}(\frak g)^L$ is a
homomorphism of Lie super algebra, is called a universal enveloping super algebra of $\frak g$ if for any other pair $(\mathcal{U}',i')$, there is an unique
homomorphism $\theta : \mathcal{U}\longrightarrow \mathcal{U}'$ such that $i' = \theta .i$. Thus, the concepts of ``super representation of $\frak g$'',
"$\frak g$-module'' and "left $\mathcal{U}(\frak g)$-module'' are completely equivalent.
Let $\frak g$ be a super Lie algebra, $\mathcal{U}(\frak g)$ be its universal enveloping super algebra, $\frak h$ be a super Lie sub-algebra of $ \frak g,$
and $V$ be
an $\frak h$-module. The $\mathbb{Z}_2$-graded space $\mathcal{U}(\frak g)\otimes _{\mathcal{U}(\frak h)} V$ can be endowed
with the structure of a $\frak g$-module as follows: $g (u\otimes v) = g u \otimes v$ for $ g \in \frak g, u\in \mathcal{U}(\frak g), v \in V.$
The so constructed $\frak g$-module is said to be induced from the $\frak h$-module $V$ and is denoted by $\mathrm{Ind}_{\frak h}^{\frak g} V$.
\subsubsection{Weights and Roots of $\frak g$.}
The standard basis for $\frak g$ consists of matrices $E_{ij}: i,j = 1\ldots, m+n$ where $E_{ij}$ is the matrix with $1$
in the place $(i,j)$ and $0$ elsewhere.
Consider the sub-algebra $\frak h$
of $\frak g$ spanned by the elements $h_j := E_{jj}: i,j = 1, \ldots ,m+n$, $\frak h$ is a Cartan subalgebra of $\frak g$.
The space $\frak h^*$ dual to $\mathcal {\frak h}$ is is spanned by $\epsilon _i : i = 1, \ldots ,m+n,$ where for $X=\left({A\ B\atop C\ D}\right)$,
$$\begin{array}{l}\epsilon _i: X\longmapsto A_{ii},\text{ for } 1\leq i\leq m, \quad
\epsilon _j:X\longmapsto D_{jj}, \text{ for } m+1\leq j\leq m+n.\end{array}$$
Elements of $\mathcal {\frak h}^*$ are called the weights of $\frak g$.
Let $\lambda \in \frak h^*$,
$\lambda =\sum_{i+1}^m \lambda _i\epsilon _i - \sum_{j= m+1}^{m+n}\lambda _j\epsilon _j$
then we write
$\lambda = ( \lambda _1,\lambda _2,\ldots, \lambda _m| \lambda _{m+1}, \ldots ,\lambda _{m+n}).$
\begin{dn}
Let $\lambda = (\lambda _1,\lambda _2,\ldots, \lambda _m| \lambda _{m+1},\ldots ,\lambda _{m+n})$ be a weight.
\begin{itemize}\item[(i)] $\lambda $ is called integral if
$\lambda _i - \lambda _{i+1} \in \mathbb{Z}$ for all $i\neq m$.
\item[(ii)] $\lambda $ is called dominant if $\lambda _i \geq \lambda _{i+1}$ for $ 1\leq i\leq m$, and
$\lambda _j\leq \lambda _{j+1}$ for $m+1\leq j\leq m+n-1$.
\item[(iii)]
$\lambda $ is called typical if $ (\lambda _i + m+1-i) - (\lambda _{m+p} + p) \neq 0$ for all $1\leq i\leq m, 1\leq p\leq n,$ otherwise it is called atypical.
\item[(iiii)]
$\lambda$ is called integrable if $\lambda_i \in \mathbb Z$ for all $i$.
\end{itemize}
\end{dn}
Let $0\neq \alpha \in \frak h^*$. Set
$ \frak g_{\alpha } := \{ a \in \frak g: [h,a] = \alpha (h) a, \forall h \in \frak h \}.$
If $\frak g_{\alpha }\neq 0$, then $\alpha$ has
the form $\epsilon_i - \epsilon _j : i \neq j$. It is called a root.
We set $\Delta _0^+ = \{ \epsilon _i - \epsilon _j : 1\leq i<j\leq m \quad \mbox{or} \quad m+1 \leq i<j\leq m+n \},
\Delta _1^+ = \{ \epsilon _i - \epsilon _j : 1\leq i\leq m \quad \mbox{and } \quad m+1 \leq j \leq m+n \} \text { and }$
$$\textstyle\rho := (m,m-1,\ldots,1|1,2,\ldots,n) - \frac{m+n+1}{2}(1,1,\ldots,1|1,1,\ldots,1).$$
\subsubsection {Kac module}
For every integral dominant weight $\lambda$, we denote by $V^0(\lambda )$ the finite dimension irreducible $\frak g_{\bar 0}$-module with highest
weight $\lambda $, $ V^0(\lambda )$ is the $(\frak g_{\bar 0} \oplus \frak g_{+1})$- module with $\frak g_{+1}$ acting by 0,
where $ \frak g_{+1}$ is the set of matrices of the form
$\left({0\ B\atop 0\ 0}\right).$
Set $\bar V(\lambda ) := \mathrm{Ind}^{\frak g}_{\frak g_{\bar 0} + \frak g_{+1}} V^0(\lambda )$. $\bar V(\lambda ) $
contains a unique maximal submodule $M(\lambda ),$ and we set
$$ V(\lambda ) : = \bar V(\lambda )/ M(\lambda ).$$
Then $V(\lambda)$ is an irreducible representation
with highest weight $\lambda$. The module $ \bar V(\lambda )$ is called generalized Verma module or Kac module
\cite{Kac3}. Kac showed that the $V(\lambda)$'s furnish all
irreducible $\frak g$-modules of finite dimension.
If $\lambda $ is typical weight then
$M_\lambda=0$, thus
$V(\lambda )=\bar V(\lambda)$, in this case $V(\lambda)$ is called typical.
On the other hand, if $\lambda$ is atypical, an explicit construction of $M(\lambda)$ is
not known.
\subsubsection{Characters of representations}
Let $V$ be a finite-dimensional irreducible $\frak g$-module. For every element $\lambda \in \mathcal{ \frak h}^*$, we define
$$V_{\lambda} := \{ v \in V :\rho(h) = \lambda(h)v \quad \mbox{for all} \quad h \in \mathcal{\frak h}\},$$
then we have $V = \oplus_{\lambda \in \mathcal{\frak h}^*}V_{\lambda}.$
The character of $V$ is
$ \mbox{\rm ch} (V) := \sum_{\lambda \in \mathcal{\frak h}^*} (\dim V_{\lambda}) e^{\lambda}.$
The following formula for the character of typical irreducible modules is due to Kac \cite{Kac3}:
\begin{equation}\label{dl0} \mbox{\rm ch} (V) = \frac{L_1}{L_0} \sum_{w \in S_m\times S_n} \mathrm{sign} (w) e^{w(\lambda + \rho )},\end{equation}
with $ L_1 = \sum _{\alpha \in \Delta _1^+}( e^{\alpha /2} + e^{-\alpha /2}), L_0 = \sum_{\beta \in \Delta _0^+} (e^{\beta /2} - e^{-\beta /2}).$
In \cite{Zhang2}, Su and Zhang gave an character formula
for all finite dimension irreducible representations with any typical and atypical dominant integral weight $\lambda $.
The formula is too complicated to recall here, but see below for a special case.
\subsection{Characters of irreducible representations of $\frak {gl}(3|1)$}
In this section, we will recall formulas for the character of all typical and atypical finite-dimensional irreducible representations of $\frak {gl}(3|1)$. According to \cite[Theorem 4.9]{Zhang2}.
In $\frak {gl}(3|1)$, we have $ \Delta _1^{+} = \{ \epsilon _1 - \epsilon _4, \epsilon _2 - \epsilon _4, \epsilon _3 -\epsilon _4 \}$,
$\Delta _0^+=\{ \epsilon _1 - \epsilon _2, \epsilon _1 - \epsilon _3, \epsilon _2 - \epsilon _3\}$. $\rho = ( \frac{1}{2}, \frac{-1}{2},
\frac{-3}{2}| \frac{-3}{2}). $
Set
$ x_1 : = e^{\epsilon _1}, x_2 := e^{\epsilon _2}, x_3 := e^{\epsilon _3}, y := e^{\epsilon _4},$
$R:= (x_1 +y)(x_2+y)(x_3+y), \Pi := (x_1-x_2)(x_2-x_3)(x_1-x_3),$
$$a (t,u,v) := \mbox{\rm det} \left( \begin{matrix}
x_1^{t+2}&x_1^{u+1}&x_1^v\\
x_2^{t+2}&x_2^{u+1}&x_2^v\\
x_3^{t+2}&x_3^{u+1}&x_3^v
\end{matrix}\right ).$$
Let $ \lambda = (\lambda _1, \lambda _2,\lambda _3|\lambda _4)$
be a typical dominant integral weight. According to the character formula (\ref{dl0}), we have
\begin{equation*
\mbox{\rm ch} (V(\lambda)) = \frac{R(x_1x_2x_3)^{\lambda _3 -1}}{\Pi y^{\lambda _4}}a(\lambda _1 - \lambda _3 ,\lambda _2 -\lambda _3 ,0).
\end{equation*}
Let $\lambda$ be an atypical weight. Then there are three possibilities:
If $\lambda _1 + 2 = \lambda _4$, then
\begin{multline}\label{ctdt2}
\mbox{\rm ch} (V(\lambda )) = \frac{R}{\Pi y^{\lambda _4}}\Big[ \frac{x_1^{\lambda _1 +2}}{x_1 +y}
( x_2^{\lambda _2}x_3^{\lambda _3-1} - x_2^{\lambda _3 -1}
x_3^{\lambda _2}) + \frac{x_2^{\lambda _1 +2}}{x_2 +y} ( x_3^{\lambda _2}x_1^{\lambda _3-1} - x_3^{\lambda _3 -1}
x_1^{\lambda _2}) \\+
\frac{x_3^{\lambda _1 + 2}}{x_3 + y}( x_1^{\lambda _2}x_2^{\lambda _3-1} - x_1^{\lambda _3 -1}
x_2^{\lambda _2}) \Big].
\end{multline}
If $\lambda _2 + 1 = \lambda _4$, then
\begin{multline}\label{ctdt3}
\mbox{\rm ch} (V(\lambda )) = \frac{R}{\Pi y^{\lambda _4}} \Big[ \frac{x_1^{\lambda _2 +1}}{x_1 +y}
( x_2^{\lambda _3-1}x_3^{\lambda _1+1} - x_2^{\lambda _1+1}
x_3^{\lambda_3-1}) + \frac{x_2^{\lambda _2 +1}}{x_2 +y} ( x_3^{\lambda _3-1}x_1^{\lambda _1+1} - x_3^{\lambda _1+1}
x_1^{\lambda _3-1}) \\+
\frac{x_3^{\lambda _2 + 1}}{x_3 + y}( x_1^{\lambda _3-1}x_2^{\lambda _1+1} - x_1^{\lambda _1+1}
x_2^{\lambda _3-1}) \Big].
\end{multline}
If $ \lambda _3 = \lambda _4$, then
\begin{multline}\label{ctdt4}
\mbox{\rm ch} (V(\lambda )) = \frac{R}{\Pi y^{\lambda _4}} \Big[ \frac{x_1^{\lambda _3}}{x_1 +y}
( x_2^{\lambda _1+1}x_3^{\lambda _2} - x_2^{\lambda _2}
x_3^{\lambda_1+1}) + \frac{x_2^{\lambda _3}}{x_2 +y} ( x_3^{\lambda _1+1}x_1^{\lambda _2} - x_3^{\lambda _2}
x_1^{\lambda _1+1}) \\+
\frac{x_3^{\lambda _3}}{x_3 + y}( x_1^{\lambda _1+1}x_2^{\lambda _2} - x_1^{\lambda _2}
x_2^{\lambda _1+1}) \Big].
\end{multline}
\section{Double Koszul complexes}
\subsection{The Koszul complex K}
In \cite{Manin1} Manin suggested the following construction to define the super determinant of a super matrix.
Let $V^*$ denote the vector space dual to $V$ with the dual basic $\xi ^1, \xi^2,\ldots ,\xi^d$, $\xi ^i(x_j) =\delta ^i_j$.
The complex $K$ has its $(k,l)$-term given by $K^{k,l} := \Lambda _k\otimes S_l^*$, where $\Lambda_k$ is the $k$-th homogeneous component of
the exterior tensor algebra over $V$, $S_l^*$ is the $l$-th homogeneous component of the symmetric tensor algebra over $V^*$.
The differential $d_{k,l}: K^{k,l} \longrightarrow K^{k+1,l+1}$ is given by
\begin{equation}\label{ct1}
d_{k,l} (h\otimes \varphi ) = \sum_i h\wedge x_i \otimes \xi ^i\cdot\varphi.
\end{equation}
In fact, the construction above gives a series of complexes $K_a$:
$$\xymatrix{K_a:\cdots\ar[r]^d&\Lambda_k\otimes S_{k-a}^*\ar[r]^d&
\Lambda_{k+1}\otimes S_{k-a+1}^*\ar[r]^d&\cdots}$$
here for $k<0$ we define $\Lambda_k$ and $S_k$ to be 0. Thus each
complex $K_a$ is bounded from below.
It is easy to check that $d_{k,l} $ is $\frak {gl}(V)$-equivariant; hence the homology groups of this complex are representations of $\frak {gl}(V)$. On
the other hand, one can show that the complex $(K_a,d)$ is exact everywhere if
$a\neq m-n$, and the complex $(K_{m-n},d)$ is exact everywhere except at the term $\Lambda_m\otimes S_n^*$,
where the homology group is one-dimensional.
This homology group defines a one-dimensional representation of $\frak {gl}(V)$. It turns out that elements of $\frak {gl}(V)$
act on this representation by means of its super determinant.
Notice that there is another differential
$\partial _{k,l} : K^{k+1,l+1}\longrightarrow K^{k,l}$, which is defined as follows:
$$\xymatrix @C=2em{\partial _{k,l}: \Lambda_{k+1}\otimes S_{l+1}\ar@{^(->}[r] &V^{\otimes k+1}\otimes V^{\otimes l+1}
\ar[rrrr]^{(\id \otimes {\mathchoice{\mbox{\rm ev} _V\otimes
\id)\circ (\id \otimes \tau _{V\otimes V^*}\otimes \id)}&&&&V^{\otimes k} \otimes V^{*\otimes l}\ar[r]^{Y_k\otimes X_l^*}&\Lambda_k\otimes S_l^*},$$
where $$ X_n := \frac {1}{n !}\sum _{w \in \sigma _n}T_w, Y_n := \frac {1}{n !}\sum_{w \in \sigma _n}(-1)^{l(w)}T_w,$$
$$ \tau (a\otimes \varphi ) = (-1)^{\hat a.\hat \varphi }\varphi \otimes a; \text { and } {\mathchoice{\mbox{\rm ev}(\varphi \otimes a) = \varphi (a), \mbox { where } a \in V, \varphi \in V^*;
a,\varphi - \mbox {homogeneous}.$$
One checks that on $ K^{k,l}$
\begin{equation}\label{ct3}
lk d_{k-1,d-1}\partial_{k-1,l-1} +(l+1)(k+1)\partial_{k,l} d_{k,l} = ( l-k -n+m ){\mathchoice{\mbox{\rm id}.
\end{equation}
Since $(K_\bullet , d)$ is exact,
$(K_\bullet , \partial )$ is also exact.
\subsection{The Koszul Complex $L$} There is another Koszul complex associated to $V$.
This complex was first defined by Priddy as a free resolution of $k$ as a module over the symmetric tensor
algebra of $V$ (see \cite{Manin}). As in the case of the complex $K$, the complex $L$ with $L^{p,r} := S_p\otimes \Lambda_r$ is defined as
a series of complexes $L_a$,
$$\xymatrix{L_a:\cdots\ar[r]^{ P}&S_p\otimes \Lambda_{a-p}\ar[r]^{ P}&
S_{p-1}\otimes \Lambda_{a-p+1}\ar[r]^{ P}&\cdots}$$
with differential $ P_{p,r}: L^{p,r} \longrightarrow L^{p-1,r+1}$
given by
$$ \xymatrix @C=4em{ P_{p,r}: S_p\otimes \Lambda_r \ar@{^(->}[r] &V^{\otimes p}\otimes V^{\otimes r}
\ar[r]^{ X_{p-1} \otimes Y_{r+1}}&S_{p-1}\otimes \Lambda_{r+1}}.$$
The complexes $(L_\bullet , P)$ are exact, except for $a = 0$.
\\
We also have another differential $Q_{p,r}: L^{p-1,r+1}\longrightarrow L^{p,r} $, given by
$$ \xymatrix@C=4em { Q_{p,r}: S_{p-1}\otimes \Lambda_{r+1}\ar @{^(->}[r]&V^{\otimes p-1}\otimes V^{\otimes r+1}= V^{\otimes p}\otimes V^{\otimes r}
\ar[r]^{\qquad \qquad X_p \otimes Y_r}& S_p\otimes \Lambda_r}.$$
One checks that on $ L^{p,r}$
\begin{equation}\label{ct60}
r (p+1) PQ + p(r+1)Q P = (p+r)\id.
\end{equation}
Consequently the complexes $(L_\bullet ,Q )$ are exact too.
\subsection{The double Koszul complex}
The main observation of this work is the fact that the two Koszul
complexes mentioned in the previous section can be combined
into a double complex which we call the double Koszul complex.
In this section we describe this complex. An application to the
construction of irreducible representations of the super Lie algebra
$\frak{gl}(3|1)$ will be given in the next section.
For simplicity we shall use the dot ``$ \cdot $'' to denote the tensor product. Fix an integer $a\geq 1$.
We arrange the Koszul
complexes $K_{-a}, K_{-a-1}, K_{-a-2}, \ldots$ as in the diagram below.
{\small $$ \xymatrix@R=1em{ K_{-a:}0\ar[r]&S^*_{a}\ar[r]^{d_{0,a}}&\Lambda_1\cdot S_{a+1}^*\ar[r]^{d_{1,a+1}}&\Lambda _2\cdot S_{a+2}^*
\ar[r]^{d_{2,a+2}}
&\Lambda _3\cdot S_{a+3}^*\ar[r]&\ldots\\
K_{-a-1}: &0\ar[r]&S_{a+1}^*\ar[r]^{d_{0,a+1}}&\Lambda_1\cdot S_{a+2}^* \ar[r]^{d_{1,a+2}}&\Lambda _2\cdot S_{a+3}^*\ar[r]&\ldots\\
K_{-a -2}: &&0\ar[r]&S_{a+2}^*\ar[r]^{d_{0,a+2}}&\Lambda_1\cdot S_{a+3}^* \ar[r]&\ldots }$$}
To get the entries on a column into a complex we tensor each complex $K_{-a-i}$ with $S_{i}$,
i.e. the complex $K_{-a-1}$ is tensored with $S_1$, the complex $K_{-a-2}$ is tensored with $S_2$, etc.
Then each column can be interpreted as the complexes
$L_j$ tensored with $S_{a+j}^*$.
Thus we have the following diagram where all rows are the Koszul complex $K_\bullet$ tensored with
$S_\bullet$ and the columns are the Koszul complex $L_\bullet$ tensored with $S_\bullet^*$:
\begin{equation}\label{phuckep0}{\small
\xymatrix@R=1em{& 0&0&0&0\\
0\ar@<.5ex>[r] &\ar[u]S_{a}^* \ar@<.5ex>[r]^d
&\ar[u]\Lambda_1\cdot S_{a+1}^* \ar@<.5ex>[r]^d &\ar[u]\Lambda _2\cdot S_{a+2}^*
\ar@<.5ex>[r]^d
&\ar[u]\Lambda _3\cdot S_{a+3}^* \ar@<.5ex>[r]^d
\ldots\\
&0\ar@<.5ex>[r] \ar@<1ex>[u] &S_1\cdot S_{a+1}^* \ar@<.5ex>[r]^d
\ar@<1ex>[u]^{ P}
&S_1\cdot \Lambda _1\cdot S_{a+2}^* \ar@<.5ex>[r]^d \ar@<1ex>[u]^{ P} &S_1\cdot \Lambda _2\cdot S_{a+3}^*
\ar@<.5ex>[r]^d
\ar@<1ex>[u]^{ P }
\ldots \\
&&0\ar@<.5ex>[r] \ar@<1ex>[u] &
S_2\cdot S_{a+2}^* \ar@<.5ex>[r]^d \ar@<1ex>[u]^{ P}
&S_2\cdot \Lambda _1\cdot S_{a+3}^*\ar@<.5ex>[r]^d
\ar@<1ex>[u]^{ P} &
\ldots \\
&&&0\ar[u]&\vdots\ar[u]}}\end{equation}
A general square in diagram (\ref{phuckep0}) has the form
\begin{equation}\label{cell}{\small\xymatrix@R=1em{
S_i\cdot \Lambda_k\cdot S_l^*\ar[r]^{{\mathchoice{\mbox{\rm id}\otimes d}& S_i\cdot \Lambda_{k+1}\cdot S^*_{l+1}\\
S_{i+1}\cdot \Lambda_{k-1}\cdot S^*_l\ar[r]_{{\mathchoice{\mbox{\rm id}\otimes d}\ar[u]^{ P\otimes{\mathchoice{\mbox{\rm id}}&
S_{i+1}\cdot \Lambda_k\cdot S_{l+1}^*\ar[u]_{ P\otimes{\mathchoice{\mbox{\rm id}} &\mbox {with} \quad l = i+k+a.\\
}}\end{equation}
For convenience, we denote $ d:= \id \otimes d, P : = P \otimes \id$.
It is easy to show that $ P d = d P$ for all above squares.
We also have an exact double Koszul complex with $d, P$ replaced by $\partial,Q$.
{\small
\begin{equation}\label{phuckep1}
\xymatrix@R=1em{& 0 \ar@{ -->}[d] &0\ar@{ -->}[d]&0\ar@{-->}[d]&0\ar@{-->}[d]
\\
0 \ar@<-.5ex>@{<--}[r]_{\partial} &S_{a}^* \ar@{ -->}[d]^Q \ar@<-.5ex>@{<--}[r]_{\partial}
& \Lambda_1\cdot S_{a+1}^* \ar@{ -->}[d]^Q\ar@<-.5ex>@{<--}[r]_{\partial} & \Lambda _2\cdot S_{a+2}^* \ar@{ -->}[d]^Q
\ar@<-.5ex>@{<--}[r]_{\partial}
& \Lambda _3\cdot S_{a+3}^*\ar@<-.5ex>@{<--}[r]_{\partial} \ar@{ -->}[d]^Q &
\ldots\\
&0\ar@<-.5ex>@{<--}[r]_{\partial} &S_1\cdot S_{a+1}^* \ar@<-.5ex>@{<--}[r]_{\partial}
\ar@{ -->}[d]^Q
&S_1\cdot \Lambda _1\cdot S_{a+2}^*\ar@<-.5ex>@{<--}[r]_{\partial} \ar@{ -->}[d]^Q &S_1\cdot \Lambda _2\cdot S_{a+3}^*
\ar@<-.5ex>@{<--}[r]_{\partial}
\ar@{ -->}[d]^Q
&
\ldots \\
&&0\ar@<-.5ex>@{<--}[r]_{\partial} &\ar@{ -->}[d]
S_2\cdot S_{a+2}^*\ar@<-.5ex>@{<--}[r]_{\partial}
&\ar@{ -->}[d] S_2\cdot \Lambda _1\cdot S_{a+3}^*\ar@<-.5ex>@{<--}[r]_{\partial}
&
\ldots\\ &&&0 &\vdots }
\end{equation}}
The commutativity of this diagram is easy to check.
\subsection{Some remarks on the structure of the double complex}
In this subsection we study some maps obtained from the differentials
of the double Koszul complex. From now, we only consider the case $(m|n) = (3|1)$.
We put the two diagrams \eqref{phuckep0} and
\eqref{phuckep1} into one:
{\small
\begin{equation}\label{phuckep3}
\xymatrix{S_{i-1}\cdot S_{a+i-1}^*\ar@<.5ex>@{^(->}[r]^{d_{0,a+i-1}} \ar@<-.5ex>@{<<--}[r]_{\partial_{0,a+i-1}}&
S_{i-1}\cdot \Lambda_1\cdot S_{a+i}^* \ar@{-->>}[d]^Q\ar@<.5ex>[r]^{d_{1,{a+i}}} \ar@<-.5ex>@{<--} [r]_{\partial_{1,{a+i}}}
&S_{i-1}\cdot \Lambda_2 \cdot S_{{a+i}+1^*}\ar@{-->}[d]^Q\ar@<.5ex> [r]^{d_{2,{a+i}+1}} \ar@<-.5ex>@{<--} [r]_{\partial_{2,{a+i}+1}}&\cdots\\
&S_i\cdot S_{a+i}^* \ar@<1ex>@{^(->}[u]^{ P}\ar@<.5ex>@{^(->}[r]^{d_{0,{a+i}}} \ar@<-.5ex>@{<<--}[r]_{\partial_{0,{a+i}}}&
S_i\cdot \Lambda _1.S_{{a+i}+1}^*
\ar@<1ex>[u]^{ P}\ar@<.5ex>[r]^{d_{1,{a+i}+1}} \ar@<-.5ex>@{<--}[r]_{\partial_{1,{a+i}+1}} \ar@{-->>}[d]^Q&
S_i\cdot \Lambda _2\cdot S_{{a+i}+2}^* \ar@{-->}[d]^Q\\
& & S_{i+1}\cdot S_{{a+i}+1}^* \ar@<1ex>@{^(->}[u]^{ P}
\ar@<.5ex>@{^(->}[r]^{d_{0,{a+i}+1}} \ar@<-.5ex>@{<<--}[r]_{\partial_{0,{a+i}+1}}\ar@<1ex>[u]^{ P}
& S_{i+1}\cdot \Lambda _1\cdot S_{{a+i}+2}^* \ar@<1ex>[u]^{ P}}
\end{equation} }
\begin{pro}\label{mde1}
The composed map $ \partial PQd : S_i\cdot S_{{a+i}}^* \longrightarrow S_i\cdot S_{{a+i}}^* $ in diagram (\ref{phuckep3}) is an isomorphism for all
$i\geq 0 $. Consequently $S_i\cdot S_{a+i}^*$ is isomorphic to a direct summand of $S_{i+1}\cdot S_{{a+i}+1}^*$.
\end{pro}
{\it Proof. }According to formulas (\ref{ct3}) and (\ref{ct60}) and the commutativity between $d,P$ and $\partial, Q$, we have
\begin{align*
\partial PQd &= \partial d - \frac{2i}{i+1}\partial Q Pd\\
& = \partial d - \frac{i}{i+1}QP + \frac {i(a+i)}{(i+1)({a+i}+1)}Qd\partial P\\
&=\Big [\frac{({a+i}+2)}{({a+i}+1)} - \frac { i}{i+1}\Big ]\id + \frac {i(a+i)}{(i+1)({a+i}+1)} Qd\partial P.
\end{align*}
We will use induction on $i$ to prove that $\partial PQd : S_{i}\cdot S_{{a+i}}^* \longrightarrow S_{i}\cdot S_{{a+i}}^*$
is diagonalizable with the set of eigenvalues
\begin{align*}\label{ctmd1}
A_i :=&\Big \{\frac{(a+i+3-j)j}{(i+1)(a+i+1)}, j=1,2,\ldots,i+1
\Big \}.
\end{align*}
For $i=0$ the claim follows from the equation above.
Assume that the proposition is true for $i-1$.
By assumption
$ \partial PQd : S_{i-1}\cdot S_{{a+i}-1}^*\longrightarrow S_{i-1}\cdot S_{{a+i}-1}^* $ is diagonalizable with the set of
eigenvalues is $A_{i-1},$
hence $Qd\partial P: S_i\cdot S_{a+i}^*\longrightarrow S_i\cdot S_{a+i}^*$ is diagonalizable with the set of eigenvalues is $A_{i-1}\cup \{ 0\}.$
Thus it is easy to see that $ \partial PQd: S_{i}\cdot S_{{a+i}}^*\longrightarrow S_{i}\cdot S_{{a+i}}^* $
is diagonalizable with $A_i$ the set of eigenvalues.\mbox{$\Box$}
Consider the diagram in (\ref{phuckep0}) as an exact sequence of horizontal complexes (except for the first column) and split it into short exact sequences.
{\small
\begin{equation}\label{phuckep4}
\xymatrix{
\ldots \ar[r] & \mbox{\rm Ker} P_{i,k}\cdot S^*_{i+k+a}\ar@<.5ex>[r]^{d'_{k,i+k+a}}\ar[d]^Q& \mbox{\rm Ker} P_{i,k+1}\cdot S^*_{i+k+a+1}\ar[d]^Q
\ar@<.5ex>[r]^{d'_{k+1,i+k+a+1}}& \mbox{\rm Ker} P_{i,k+2}\cdot S^*_{i+k+a+2} \ar[r]\ar[d]^Q
&\ldots \\
\ldots\ar[r]&S_{i+1}\cdot \Lambda _{k-1} \cdot S^*_{i+k+a}\ar[d]^Q\ar@<1ex>@{->>}[u]^{ P_{i+1,k-1}}
\ar@<.5ex>[r]^{d_{k-1,i+k+a}}
&S_{i+1}\cdot \Lambda _k\cdot S^*_{i+k+a+1}\ar[d]^Q
\ar@<1ex>@{->>}[u]^{ P_{i+1,k}} \ar@<.5ex>[r]^{d_{k,i+k+a+1}}
&S_{i+1}\cdot \Lambda _{k+1}\cdot S^*_{i+k+a+2}\ar@<1ex>@{->>}[u]^{ P_{i+1,k+1}} \ar[r]\ar[d]^Q&\ldots\\
\ldots\ar[r]&\mbox{\rm Ker} P_{i+1,k-1}\cdot S^*_{i+k+a} \ar@<.5ex>[r]^{d'_{k-1,i+k+a}} \ar@<1ex> @{^(->}[u]^{i}&\mbox{\rm Ker} P_{i+1,k }
\cdot S_{i+k+a+1}^*
\ar@<.5ex>[r]^{d'_{k,i+k+a+1}} \ar@<1ex> @{^(->}[u]^{i}&
\mbox{\rm Ker} P_{i+1,k+1}\cdot S^*_{i+k+a+2}\ar@<1ex>@{^(->}[u]^{i}\ar[r]&\ldots }
\end{equation}}
where $d'_{k,i+k+a} := d_{k,i+k+a}\big |_{\mbox{\rm Ker} P_{i,k}\cdot S_{i+k+a}^*}$ , $\mbox{\rm Ker} P_{i,j} = \mbox{\rm Im} P_{i+1,j-1}$ for all $i\geq 0.$
The differentials $\partial$ however do not restrict to differentials on the first and the
third horizontal complexes.
Consider the following part of \eqref{phuckep4} for $i, k\geq 1$:
{\small
\begin{equation}\label{sodo}\xymatrix{ S_{i-1}\cdot \Lambda_{k+1}\cdot S_{a+i+k}^*\ar[r] \ar @<-.5 ex>@{-->}[d]
&S_{i-1}\cdot \Lambda_{k+2}\cdot S_{a+i+k+1}\ar@<-.5ex>@{-->}[l] \ar @<-.5 ex>@{-->}[d] &\\
\mbox{\rm Ker} P_{i-,k+1}\cdot S^*_{a+i+k}\ar[r] \ar @{^(->}[u] \ar @<-.5ex>@{-->}[d]_Q&
\mbox{\rm Ker} P_{i-1,k+2}\cdot S_{a+i+k+1}^*\ar @{^(->}[u] \ar @<-.5ex>@{-->}[d] \ar@<-.5ex>@{-->}[l] &\\
S_i\cdot \Lambda_k \cdot S_{a+i+k}^*\ar[r]_d \ar @{->>} [u]_P &\ar@<-.5ex>@{-->}[l]_\partial \ar @{->>}[u]S_i\cdot \Lambda_{k+1} \cdot S_{a+i+k+1}^*\ar[r] \ar @<-.5 ex>@{-->}[d]& \cdot S_i \cdot \Lambda_{k+2}\cdot S^*_{a+i+k+2}
\ar@<-.5ex>@{-->}[l] \ar @<-.5 ex>@{-->}[d] \\
& \mbox{\rm Ker} P_{i,k+1}\cdot S^*_{a+i+k+1}\ar[r] \ar @{^(->}[u] \ar @<-.5ex>@{-->}[d]_Q&
\mbox{\rm Ker} P_{i,k+2}\cdot S_{a+i+k+2}^*\ar @{^(->}[u] \ar @<-.5ex>@{-->}[d] \ar@<-.5ex>@{-->}[l] \\
& S_{i+1}\cdot \Lambda_{k}\cdot S^*_{a+i+k+1}\ar[r]_d \ar @{->>}[u]_{ P}&
S_{i+1}\cdot \Lambda_{k+1}\cdot S_{a+i+k+2}^* \ar @{->>}[u] \ar@<-.5ex>@{-->}[l]_{\partial} }
\end{equation} }
\begin{pro}\label{dl1}
The composed map $$ P\partial dQ: \mbox{\rm Ker} P_{i,k+1}\cdot S_{a+i +k+1}^*\longrightarrow \mbox{\rm Ker} P_{i,k+1}\cdot S_{a+i+k+1}^*$$
(for $i\geq 0, k\geq 1$) in the diagram (\ref{sodo})
is an isomorphism. Consequently $\mbox{\rm Ker} P_{i,k+1}\cdot S_{a+i+k+1}^*$ is isomorphic to a direct summand of $S_{i+1}\cdot \mbox{\rm Im} d_{k,a+i+k+1}.$
\end{pro}
{\it Proof.}
By using the method of induction,
we will prove that $$ P\partial dQ: \mbox{\rm Ker} P_{i,k+1}\cdot S_{a+i +k+1}^*\longrightarrow \mbox{\rm Ker} P_{i,k+1}\cdot S_{a+i+k+1}^*,$$
is diagonalizable with the set of eigenvalues is
{\small
\begin{align*}
&A_i := \Big \{\frac{(a+k+2i+4-j)j}{(i+1)(k+1)^2(a+i+k+2)}, j=1,2,\ldots,i+1,i+k+1
\Big \}.
\end{align*}}
For $i =0$, consider the following part of \eqref{sodo}:
{\small
\begin{equation*} \xymatrix{\Lambda_{k }\cdot S_{a+k }^*\ar[r]_d \ar @<-.5 ex>@{-->}[d]_Q& \ar@<-.5ex>@{-->}[l]_{\partial}
\Lambda_{k+1}\cdot S_{a+k+1}^*\ar[r]_d \ar@<-.5ex>@{-->}[d]_Q&
\cdot \Lambda_{k+2}\cdot S^*_{a+k+2}\ar@<-.5ex>@{-->}[d] \ar@<-.5ex>@{-->}[l]_{\partial} \\
\cdots S_1\cdot\Lambda_{k-1}\cdot S^*_{a+k}\ar[r]_d \ar@{->>}[u]_{ P} & S_{1}\cdot \Lambda_{k}\cdot S^*_{a+k+1}\ar[r]_d \ar@{->>}[u]_{ P}\ar@<-.5ex>@{-->}[l]_{\partial}&
S_{1}\cdot \Lambda_{k+1}\cdot S_{a+k+2}^*\ar@{->>}[u] \ar@<-.5ex>@{-->}[l]_{\partial}
}\end{equation*} }
The composed map $ P\partial DQ: \Lambda_{k+1}\cdot S_{a+k+1}^*\longrightarrow \Lambda_{k+1}\cdot S_{a+k+1}^*.$
By means of formulas (\ref{ct3}) and (\ref{ct60}) we have
\begin{align*}
P\partial dQ = P\frac{[(a+3)- k(a+k+1)d\partial ]}{(k+1)(a+k+2)} Q = \frac {(a+3)} {(k+1)(a+k+2)}\id - \frac {k(a+k+1)}{(k+1)(a+k+2)}d \partial.\\
\end{align*}
We have $d\partial $ is diagonalizable with eigenvalues $0$ and $ \frac {a+2}{(k+1)(a+k+1)}$, hence
$ P\partial dQ$ is diagonalizable with the set of eigenvalues
$$ A_0 := \Big \{\frac {(a+3)}{(k+1)(a+k+2)},
\frac {(a+k+3)}{(k+1)^2(a+k+2)}\Big \}.$$
For $i=1$,
consider in \eqref{sodo} the map $$ P\partial dQ: \mbox{\rm Ker} P_{1,k+1}\cdot S_{a+k+2}^*\longrightarrow \mbox{\rm Ker} P_{1,k+1}\cdot S_{a+k+2}^*.$$
On $S_i\cdot \Lambda_{k+1}\cdot S^*_{a+i+k+1}$, we have
{\small
\begin{align*}
P\partial dQ &= P\frac{[(a+4) - k(a+k+2)d\partial ]}{(k+1)(a+k+3)}Q = \frac {(a+4)}{(k+1)(a+k+3)} PQ -\frac{k(a+k+2)}{(k+1)(a+k+3)} P d\partial Q. \end{align*}}
We have $Pd\partial Q=dPQ\partial$ and this operator can be restricted
to $\mbox{\rm Ker} P_{1,k+1}\cdot S_{a+k+2}^*$. We compute the eigenvalue of
this operator. First we have
\begin{align*}
dPQ\partial&= d[\frac{(k+1) - (k+1) QP}{2 k}]\partial
= \frac { k+1 }{2 k }d\partial - \frac { k+1 }{2 k }dQP\partial.
\end{align*}
Notice that $dQP\partial$ is an endomorphism of $\mbox{\rm Im} d\subset S_i\cdot \Lambda_{k+1}\cdot S^*_{a+i+k+1}$, on this space $d\partial$
operates by multiplication with $ \frac {a+3}{(k+1)(a+k+2)}$.
On the other hand from above we know the eigenvalues of
$P\partial dQ$ form the set $A_0$. Thus
$dQP\partial$ is diagonalizable with eigenvalues $A_0\cup \{0 \}$.
Consequently $dPQ\partial$ is diagonalizable with eigenvalues
$$\Big\{\frac{a+3}{2k(a+k+2)}, \frac{a+2}{2(k+1)(a+k+2)},0\Big\}.$$
On the other hand, the restriction of $PQ$ to $\mbox{\rm Ker} P_{1,k+1}\cdot S_{a+k+2}^*$ is the multiplication with $\frac{k+2}{2(k+1)}$.
Therefore, the eigenvalues of $P\partial dQ$ are
\begin{align*}
A_1 := & \Big \{\frac {(a+4)(k+2)}{2(k+1)^2(a+k+3)}, \frac{(a+k+5)}{2(k+1)^2(a+k+3)}, \frac {2(a+k+4)}{2(k+1)^2(a+k+3)}
\Big \}.
\end{align*}
In general,
we consider the composed map $$P\partial dQ: \mbox{\rm Ker} P_{i,k+1}\cdot S_{a+i+k+1}^* \longrightarrow \mbox{\rm Ker} P_{i,k+1}\cdot S_{a+i+k+1}^*
\mbox { in the diagram (\ref{sodo}).}$$
We have
{\small
\begin{align*}
&P\partial dQ = P[\frac{(a+i+3) - k(a+i+k+1)d\partial }{(k+1)(a+i+k+2)}]Q\\
&= \frac {(a+i+3)}{(k+1)(a+i+k+2)}PQ- \frac{k(a+i+k+1)}{(k+1)(a+i+k+2)}dPQ\partial \\
&= \frac{(a+i+3)(i+k+1)}{(k+1)^2(i+1)(a+i+k+2)}\id -\frac{k(a+i+k+1)}{(k+1)(a+i+k+2)}d[\frac {(i+k) - i(k+1)QP}{k(i+1)}]\partial \\
&= \frac{(a+i+3)(i+k+1)}{(k+1)^2(i+1)(a+i+k+2)}\id - \frac{(i+k)(a+i+k+1)}{(k+1)(i+1)(a+i+k+2)}d\partial
+ \frac{i(a+i+k+1)}{(i+1)(a+i+k+2)}dQP\partial .
\end{align*}}
Similar arguments shows that $dQP\partial: \mbox{\rm Ker} P_{i,k+1}\cdot S_{a+i+k+1}^* \longrightarrow
\mbox{\rm Ker} P_{i,k+1}\cdot S_{a+i+k+1}^*$ is diagonalizable with the set of eigenvalues being $A_{i-1}\cup \{0\}$. Thus
the composed map $ P\partial dQ: \mbox{\rm Ker} P_{i,k+1}\cdot S_{a+i+k+1}^* \longrightarrow
\mbox{\rm Ker} P_{i,k+1}\cdot S_{a+i+k+1}^*$ is diagonalizable with $A_i$ the set of eigenvalues, hence is an isomorphism.\mbox{$\Box$}
\section{Construction of irreducible representations of $\frak {gl}(V)$.}
Let $V$ be a super vector space with super-dimension $(3|1)$.
In this section, using the double Koszul complex, we will construct all irreducible representations of this super algebra. To show the representations
obtained are in fact irreducible we compute their characters.
\subsection{Combinatorial construction of irreducible representations of $\frak {gl}(V)$}
In this section, we will compute the character of the duals of irreducible direct summand of the power of the
fundamental representation $V$.
By the combinatorial method, we have
$$V^{\otimes k} = \bigoplus_{\lambda \in \Gamma_{3,1} } I_{\lambda }^{\oplus C_{\lambda }},$$
where $ I_{\lambda }$ are simple, and
$\Gamma_{3,1} $ is the set of partitions with
$\lambda _4 \leq 1.$
Since the character of $V$ is $x_1+x_2+x_3-y$, using the determinant formula (3.5) of \cite{Macdonald}, we can compute the
character of $I_\lambda $ for all $ \lambda \in \Gamma_{3,1}$.
If $\lambda \in \Gamma _{3,1}$ and $\lambda_3\geq 1$, we have
\begin{equation}\label{ctdt1}
\mbox{\rm ch} (I_{\lambda _1,\lambda _2,\lambda _3,1^\lambda _4} ) = \frac{R(x_1x_2x_3)^{\lambda _3 -1}}{\Pi y^{\lambda _4}}
a(\lambda _1 - \lambda _3 ,\lambda _2 -\lambda _3 ,0),
\end{equation}
hence
\begin{equation*}
\mbox{\rm ch} (I_{\lambda _1,\lambda_2,\lambda_3,1^\lambda_4}^*)= \frac{R (x_1x_2x_3)^{-\lambda_1}}{\Pi y^{\lambda_4+3}}a(\lambda _1-\lambda _3 ,
\lambda _1-\lambda _2 ,0).
\end{equation*}Thus
$I_{\lambda _1,\lambda_2,\lambda_3,1^\lambda_4}^*$ has highest weight $(-\lambda _3+1,-\lambda _2+1,-\lambda _1 +1|\lambda_4+3).$
Therefore we have $$\mbox{\rm ch} (I_{\lambda _1,\lambda _2,\lambda _3,1^\lambda _4} ) = \mbox{\rm ch} (V(\lambda _1,\lambda _2,\lambda _3|-\lambda _4)), $$
$$\mbox{\rm ch} (I_{\lambda _1,\lambda_2,\lambda_3,1^\lambda_4}^*) = \mbox{\rm ch} (V(-\lambda _3+1,-\lambda _2+1,-\lambda _1 +1|\lambda_4+3)).$$
Now,
\begin{align*}
\mbox{\rm ch} (I_{\lambda_1,\lambda_2,0,0}) &= \frac{R}{\Pi }\Big [ \frac{x_2^{\lambda_1+1}x_3^{\lambda_2} - x_2^{\lambda_2} x_3^{\lambda_1+1}}
{x_1 + y}
+ \frac{x_3^{\lambda_1+1}x_1^{\lambda_2} - x_3^{\lambda_2} x_1^{\lambda_1+1}}{x_2 + y}
+ \frac{x_1^{\lambda_1+1}x_2^{\lambda_2} - x_1^{\lambda_2} x_2^{\lambda_1+1}}{x_3 + y} \Big],
\end{align*}
hence
\begin{multline*}
\mbox{\rm ch} (I_{\lambda_1,\lambda_2,0,0}^*) = \frac{R}{\Pi y^2} \Big[ \frac{x_1^2}{x_1 + y}(x_2^{-\lambda_2+1}x_3^{-\lambda_1} - x_2^{-\lambda_1}
x_3^{-\lambda_2+1}) +
\frac{x_2^2}{x_2 + y}(x_3^{-\lambda_2+1}x_1^{-\lambda_1} - x_3^{-\lambda_1}x_1^{-\lambda_2+1}) \\
+ \frac{x_3^2}{x_1 + y}(x_1^{-\lambda_2+1}x_2^{-\lambda_1} - x_1^{-\lambda_1}x_2^{-\lambda_2+1}) \Big].
\end{multline*}
Therefore $I_{\lambda_1,\lambda_2,0,0}^*$ has highest weight
$(0,-\lambda _2+1,-\lambda _1+1|2).$
Thus we have
$$ \mbox{\rm ch} (I_{\lambda_1,\lambda_2,0,0}) = \mbox{\rm ch} (V(\lambda_1,\lambda_2,0|0)),$$
$$ \mbox{\rm ch} (I_{\lambda_1,\lambda_2,0,0}^*) = \mbox{\rm ch} (V(0,-\lambda _2+1,-\lambda _1+1|2)).$$
Further we have
\begin{equation}\label{ctdt8}
\mbox{\rm ch} (I_{\lambda _1,0,0,0}) = \frac{1}{\Pi } \Big[ x_2^{\lambda _1+1}(x_2+y)(x_3-x_1) + x_3^{\lambda _1+1}(x_3+y)(x_1-x_2) +
x_1^{\lambda _1+1}(x_1+y)(x_2-x_3) \Big],
\end{equation}and hence
\begin{align*}
\mbox{\rm ch} (I_{\lambda _1,0,0,0}^*)
&= \frac{1}{\Pi y}\Big [ x_1^2(-x_2^{-\lambda _1+1}x_3 + x_2 x_3^{-\lambda _1+1}) + x_2^2(-x_3^{-\lambda _1+1}x_1
+ x_3 x_1^{-\lambda _1+1})\\
&+
x_3^2(-x_1^{-\lambda _1+1}x_2
+ x_1 x_2^{-\lambda _1+1})
+ x_1^2y(-x_2^{-\lambda _1}x_3 +x_2 x_3^{-\lambda _1}) \\
&+ x^2y(-x_3^{-\lambda _1}x_1 + x_3 x_1^{-\lambda _1})
+
x_3^2y(-x_1^{-\lambda _1}x_2 + x_1 x_2^{-\lambda _1}) \Big ].
\end{align*}
Therefore $I_{\lambda_1,0,0,0}^*$ has highest weight $(0,0,-\lambda _1 +1|1)$.
Thus we have
$$\mbox{\rm ch} (I_{\lambda_1,0,0,0}) = \mbox{\rm ch} (V(\lambda_1,0,0|0)), $$
$$\mbox{\rm ch} (I_{\lambda_1,0,0,0}^*) = \mbox{\rm ch} (V(0,0,-\lambda _1 +1|1)).$$
\subsection{Construct representations by using Koszul complex $K$}
Consider complexes $K_a$, with $a: = k-l \neq 2$.
$$K_a: ....\longrightarrow \Lambda _k.S_l ^*\longrightarrow \Lambda _{k+1}.S_{l+1}^*\longrightarrow \Lambda _{k+2}.S_{l+2}^*\longrightarrow ....,$$
By using the exactness property of the Koszul complex $K$, we will construct a class of irreducible representations of $\frak {gl}(3|1)$. According to (\ref{ct3}) we have
\begin{equation}\label{xdanh}
\Lambda _k.S_l^* \cong \mbox{\rm Im} d_{k-1,l-1} \oplus \mbox{\rm Im} d_{k,l}.
\end{equation}
Consequently, we have (\cite{dh2}).
\begin{pro}
The module $\mbox{\rm Im} d_{k+1,l+1}$ is simple for all pairs $(k,l)$ with $l,k \geq 1, k-l \neq 2.$
\end{pro}
Using induction, we find that
\begin{equation}\label{ctdt7}
\mbox{\rm ch} (\mbox{\rm Im} d_{k,l}) = \frac{Ry^{k-3}}{\Pi (x_1x_2x_3)^l}a (l ,l ,0).
\end{equation}
Set $M^{m,p} := \mbox{\rm Im} d_{m+2,m+p} . H_{3,1}^{\otimes m-1}$ with $H_{3,1}:= \mbox{\rm Ker} d_{3,1}/\mbox{\rm Im} d_{2,0}$. We have
\begin{equation*}
\mbox{\rm ch} (M^{m,p} )= \frac{R}{\Pi (x_1x_2x_3)^{p+1}}a (m+p , m+p , 0) = \mbox{\rm ch} (V(m,m,-p|0)).
\end{equation*}
Hence
\begin{equation*}
\mbox{\rm ch} (M^{m,p })^* =\frac {R (x_1x_2x_3)^{-m}}{\Pi y^3}a(m+p , 0,0)= \mbox{\rm ch} V(p+1,-m+1,-m+1|3).
\end{equation*}
Therefore, $M$ is isomorphic to $V(m,m,-p|0) $, and $M^*$ is isomorphic to $V(p+1,-m+1,-m+1|3). $
Thus every irreducible representation with highest weight in the set
$$\{ (m,n,p|-q), (-p,-n,-m|q): (m,n,p,q) \in \Gamma _{3,1}\}\cup
\{(m,m,-p|0) , (p,-m,-m|0): m,p\geq 1\} $$
is constructed.
It remains to construct representations with highest weights in the set
$$\{(n,0,-p|0): n,p\geq 1\} \cup \{ (m+a,m, -p|,0): m,a,p\geq 1\}.$$
\subsection{Construct representations by using double Koszul complex.}
According to Prop. \ref {mde1}, there exists $Y$ such that
$ S_{n}\cdot S_{p}^* = S_{n-1}\cdot S_{p-1}^* \oplus Y$. It is easy to compute for $n,p\geq 1$ that
\begin{equation}\label{ctdt11}
\mbox{\rm ch} (Y )= \frac{(x_1x_2x_3)R}{\Pi y}\Big[ \frac{x_2^{-p-1}x_3^n - x_2^nx_3^{-p-1}}{x_1+y} +
\frac{x_3^{-p-1}x_1^n - x_3^nx_1^{-p-1}}{x_2+y} + \frac{x_1^{-p-1}x_2^n - x_1^nx_2^{-p-1}}{x_3+y} \Big ].
\end{equation}
Hence, $Y$ has highest weight
$ (n,0,-p+1|1).$
Next, we will construct representations having highest weights in the set $\{ \lambda = (m+t,m,-p,0): m,p,t \geq 1\}.$
According to Proposition \ref{dl1}, we have $$ S_1\cdot \mbox{\rm Im} d_{2,m+1} = \Lambda _3\cdot S_{m+1}^* \oplus Z_1,$$ hence
\begin{equation}\label{ctdt22}
\mbox{\rm ch} (Z_1 )= \frac{R}{\Pi y(x_1x_2x_3)^{m+1}}a(m+2,m+1,0)= \mbox{\rm ch} (V(2,1,-m+1|1)),
\end{equation}
therefore $Z_1$ is isomorphic to $V(2,1,-m+1|1).
In general, according to Proposition \ref {dl1}, we have
$\mbox{\rm Im} (\id_{I_{k,0,0,0}}\cdot d_{l,m}) = \mbox{\rm Ker} P_{k,l}\cdot S_m^* \oplus Z_k,$ where $\mbox{\rm Ker} P_{k,l} \cong I_{k,1^l}$. Therefore
$\mbox{\rm ch} (Z_k) = \mbox{\rm ch} [\mbox{\rm Im} (\id_{I_{k,0,0,0}}\cdot d_{l,m})] - \mbox{\rm ch} (I_{k,1^l}\cdot S_m^*).$
According to (\ref{ctdt1}), (\ref{ctdt8}) and (\ref{ctdt7}), we have
\begin{equation}\label{ctdt24}
\mbox{\rm ch} Z_k= \frac{R(x_1x_2x_3)^{-m}y^{l-3}}{\Pi }a(k+m,m-1,0).
\end{equation} Set $M := Z_k\cdot I_{1,1,1,-1}^{\otimes (l-2)}$, then
\begin{equation}\label{ctdt25}
\mbox{\rm ch} (M )= \frac{R(x_1x_2x_3)^{-p}}{\Pi y}a(m+p+t-1, m+p-1, 0) = \mbox{\rm ch} (V(m+t,m,-p+1|1)),
\end{equation} where $t :=l-1,
p := -m-2+l.$
According to (\ref{ctdt25}), we have
\begin{equation}\label{ctdt26}
\mbox{\rm ch} (M^*)= \frac{R(x_1x_2x_3)^{-m-a}}{\Pi y^2}a(m+t+p-1, t,0)=\mbox{\rm ch} (V(p+1,-m+1,-m-t+1|3)).
\end{equation}
Therefore $M$ is isomorphic to $ V(m+t,m,-p+1|1)$, $M^*$ is isomorphic to $V(p+1,-m+1,-m-t+1|3). $
Thus, for any integrable dominant weight $\lambda = (\lambda_1, \lambda_2, \lambda_3| \lambda_4 ) $, we have constructed a representation
which has highest weight $ \lambda $ and has character equal to the character of the irreducible representation with highest weight $\lambda$.
|
\section{Introduction}
Prominences protruding out of the perfect sphere of the visible solar disk are even
visible with the naked eye, when the bright disk is occulted. These enigmatic
features, which apparently withstand gravity, have attracted scientists since centuries, but despite of substantial
progress, which has been made during the last decades in understanding the
physics of prominences, important aspects are still not understood. The enormous
efforts to understand the nature of prominences is reflected in a wealth of
literature. We refer to review articles and reference material, which may be relevant
to our work \citep[e.g., ][]{Tandberg95,Spiros02,Parenti05,Wilhelm07}, and focus
on prominence observations of the hydrogen Ly$-\alpha$~ line profile,
which reveal information on the physical conditions for the line formation.
Anywhere on the disk, this line is self-reversed \citep{Curdt01}.
The reversal in the profile is, among others, related to the amount of neutral
hydrogen in its ground level, which by itself is a complex function of the
temperature and density structure of the emitting plasma. In addition,
flows of the emitting or the absorbing plasma and magnetic field may modulate
the sizes of the red or the blue peak and the symmetry of the profile \citep{Curdt08,Tian09a}.
Early observations of the Ly$-\alpha$~ line profile in prominences were completed with
the LPSP instrument on {\it OSO 8} \citep{Vial82} and the UVSP instrument on
{\it SMM} \citep{Fontenla88}. These photoelectric measurements had to be corrected
for the geocoronal absorption. They already have shown signatures of asymmetry
and a wide parameter range for the depth of the reversal of the profile,
features which at that time could not be reproduced by radiative transfer calculations.
Later on, the modeling work made it clear that the overall emergent profile
highly depends on the physical conditions in the prominence \citep[e.g.,][]{Gouttebroze93}.
In particular, the imprint of the incident profile and the role of the
Prominence Corona Transition Region (PCTR) were now employed to reproduce
observations \citep[e.g.,][]{Vial07}.
Recently, 2D-multithread models were established, which are based on theoretical work of
\citet{Heinzel01} and predict -- depending on the orientation of the prominence axis
relative to the line of sight (LOS) -- opposite asymmetries of the Ly$-\alpha$~ and Ly$-\beta$~ lines \citep{Gunar07,Gunar08}
and deeply or less deeply self-reversed profiles \citep{Schmieder07}.
This is the dedicated context and the rationale of our work.
\section{Observations}
Because of its wavelength range from 660~{\AA} to 1600~{\AA}, its high spectral
resolution, and its vantage point outside of the irritating geocorona,
which absorbs Ly$-\alpha$~ emission, the SUMER instrument on {\it SOHO} \citep{Wilhelm95}
is ideally suited to provide information about the line profile.
Its enormous brightness exceeds however the capabilities
of the SUMER detectors and Ly$-\alpha$~ can only be observed in small sections of 50 px
on both sides of the detectors beneath a 1:10 attenuating grid.
Unfortunately, the attenuation also exerts a modulation onto
the line profile, which makes it difficult to interpret such data. Attempts
made to observe Ly$-\alpha$~ in quiet Sun locations on the unattenuated bare section
of the photocathode had difficulties to calibrate the local gain depression.
First results from prominence data acquired in spring 2005 have
been reported by \citet{Gunar06}, \citet{Vial07}, and \citet{Gunar08}.
In July 2008, the SUMER team found a new, unconventional method to observe the extremely
bright Ly$-\alpha$~ line of neutral hydrogen with partially closed telescope aperture to
reduce the incoming photon flux. The obtained genuine Ly$-\alpha$~ profiles in the quiet
Sun and coronal hole regions were analysed by \citet{Curdt08}, \citet{Tian09a}, and
\citet{Tian09b}. Here we present, for the first time, unprecedented Ly$-\alpha$~ observations of
two quiescent prominences seen in June 2009 and discuss the results obtained from a
detailed analysis of the line profiles.
\begin{figure}
\includegraphics[width=8cm]{f1_13875.eps}
\caption{The prominences observed above the southeast limb on June 15 (left)
and above the southwest limb on June 9 (right). The images are taken in the
EIT 304 channel \citep{Boudin95} and the area of the SUMER rasters is indicated by rectangles.
}
\label{scanregion}
\end{figure}
\begin{figure*}
\sidecaption
\includegraphics[width=12cm]{f2_13875.eps}
\caption{Raster scan in Ly$-\alpha$~ and Si\,{\sc iii} of the prominence observed
on June 15. The line radiance is given in counts/px/s.
The raster also covers a small disk section near the southeast limb.
Solid and dotted contours represent top 40\,\% and top 15\,\%, respectively.
The Ly$-\alpha$~ contours have been transferred to the Si\,{\sc iii} raster.
We show the average profiles of Ly$-\alpha$~ of six distinguished locations (again in instrumental units):
\vspace{2mm}
\newline
(1) disk,\newline
(2) limb and near disk,\newline
(3) sub-prominence void,\newline
(4) inner prominence boundary,\newline
(5) prominence core,\newline
(6) outer prominence boundary.
}
\label{Jun15}
\end{figure*}
\begin{figure*}
\sidecaption
\includegraphics[width=12cm]{f3_13875.eps}
\caption{Idem for the observation on June 9. The raster was completed just outside the southwest limb.
We show the average profiles of Ly$-\alpha$~ of six distinguished
locations:
\vspace{2mm}\newline
(1) inner prominence core\newline
(2) prominence interconnection\newline
(3) sub-prominence void\newline
(4) prominence interconnection\newline
(5) outer prominence core\newline
(6) outer prominence boundary.
}
\label{Jun9}
\end{figure*}
The new method to reduce the incoming photon flux to a moderate level,
appropriate for Ly$-\alpha$, was described in earlier work \citep{Curdt08,Tian09a,Tian09b}.
A standard procedure to partially close the door led to a reproducible
reduction to a 20\,\% level.
In June 2009, this method was applied for the first time to prominence observations.
On June 9 and 15, 2009, we completed raster scans of approximately 22\hbox{$^{\prime\prime}$}~$\times$~120{\hbox{$^{\prime\prime}$}}
at positions near the solar limb with mid-size prominences.
Two spectral windows were transmitted, 100~pixels (px) around Ly$-\alpha$ recorded on the bare
photocathode of the detector, and 50~px around $\lambda$\,1206~Si\,{\sc iii} recorded
on the KBr-coated section of the photocathode, respectively. All observations were completed with
the 0.3\hbox{$^{\prime\prime}$}~$\times$~120{\hbox{$^{\prime\prime}$}} slit. With an exposure time of 14.5~s, both lines
were observed with sufficient counts for a good line profile analysis.
For both data sets three exposures at each position were completed before the
raster was continued with a very small increment of 0.375{\hbox{$^{\prime\prime}$}}.
From a first inspection, we found no indication for temporal variations of the
object during the observing time of 45 minutes. In our statistical analysis
we keep the temporal information and assume that the sub-resolution
increment of 0.375{\hbox{$^{\prime\prime}$}} (= 3/8{\hbox{$^{\prime\prime}$}}) is equivalent to three
hypothetical increments of 1/8{\hbox{$^{\prime\prime}$}}.
The exact knowledge of the limb position and the limb distance for each pixel
is very important for prominence observations. Therefore we used additional
information, for an independent assessment of the instrument pointing uncertainty,
provided by the hardware encoders in the instrument's housekeeping channel.
Thus, we confirmed that the azimuth movement was as expected and that the
actual east-west pointing was very close to its nominal value.
Similarly, we confirmed that in elevation the absolute positions for both
rasters differ by the nominal value of 5{\hbox{$^{\prime\prime}$}}. Since in the June 9 data set
the position of the limb can be determined, we can estimate that the overall
pointing uncertainty is in the order of 2{\hbox{$^{\prime\prime}$}} to 3{\hbox{$^{\prime\prime}$}}. The prominences as seen
in the EIT 304 channel are shown in Fig.\,\ref{scanregion},
white rectangles marking the area covered by the SUMER rasters.
Both data sets were processed with standard procedures of the
SUMER-soft library. We used the dedicated flatfield exposure of April~19, 2009
to complete the flatfield correction.
\section{Prominences in Ly$-\alpha$~ and in Si\,{\sc iii}}
The rasters for both days are shown in Fig.\,\ref{Jun15} and Fig.\,\ref{Jun9}.
Note, that the x-axis also contains time information (cf., prev. section)
and that both axes are at a different scale, and that the x-dimension is stretched.
The contours delineate the top 15\,\% (dashed) and the top 40\,\% (solid) of the
pixels in the Ly$-\alpha$~ brightness histogram. These contours have been transferred
to the Si\,{\sc iii} raster,
In Fig.\,\ref{Jun15} (observation on June~15) we distinguish six different segments of the raster,
separated by the blue contours or red boxes:\\
(1) disk,
(2) limb and near disk,
(3) sub-prominence void,\\
(4) inner prominence boundary,
(5) prominence core,\\
(6) outer prominence boundary.
We also display the averaged profiles of the designated areas. It is obvious
that the disk profiles are much wider and more reversed than the prominence
profiles. The disk profiles are almost symmetric, which is consistent with
the downflow argument in \citet{Curdt08} for this special geometry.
Interestingly, a clear blue-peak dominance is observed in the profile of the
prominence core.
In Fig.\,\ref{Jun9} (observation on June~9) we distinguish:\\
(1) inner prominence core
(2) prominence interconnection\\
(3) sub-prominence void\hspace{2mm}
(4) prominence interconnection\\
(5) outer prominence core
(6) outer prominence boundary.
Here, we observe just outside the limb. The contrast is even lower in this
prominence, which is shown with the same dynamic range. The profiles are
almost flat-topped, significant reversals are not seen anywhere. The blue-peak
dominance is also present in both parts of this prominence, but less evident.
The prominence is also seen in Si\,{\sc iii}. Again, the Ly$-\alpha$~ radiance contours
have been transferred. These contours show, that there are considerable differences in
the Si\,{\sc iii} spectroheliogram; structures are not co-spatial; the
prominence appears more granulated and not as diffuse as in Ly$-\alpha$. The
formation temperature of Si\,{\sc iii} is 70\,000~K, much higher than typical
prominence temperatures of 6\,000~K to 8\,000~K. Si\,{\sc iii} is a typical
transition region line. Since its wavelength is well
above 912~{\AA}, the Lyman limit, opacity effects by hydrogen can be ruled out.
The prominence is basically transparent \citep{Anzer07}. These authors also show
that the C\,{\sc i} recombination continuum below 1239~{\AA} is negligible
and, consequently, the PCTR of each unresolved thread would contribute to the
Si\,{\sc iii} emission and one would expect an appearance similar to the
cold body. The differences in appearance may be an indication for the coexistence
of hot and cold plasma with different opacities. Recent observations by {\it Hinode}-SOT
\citep{Berger08} assume buoyant bubbles of hotter plasma in quiescent prominences,
although on smaller scales. Such a scenario would also be compatible with
our observation. Without {\it Hinode}-SOT co-observations, however, our results remain
inconclusive.
We sorted the pixels of all disk locations and of all prominence locations by the
total line radiance and defined six equally spaced radiance bins.
The profiles for these bins are displayed in Figs.\,\ref{ProfJun15} and \ref{ProfJun9}.
There are striking differences of prominence profiles compared to disk profiles.
In the prominence, the contrast is much lower, reduced by a factor of 4 to 5.
The blue-peak dominance is observed in all radiance bins from the brightest
areas of the prominence core.
\begin{figure}
\includegraphics[width=8cm]{f4_13875.eps}
\caption{We have sorted the pixels within the top 40\,\% contours (solid in Fig.\,\ref{Jun15}) by their
radiance and show the profiles of Ly$-\alpha$~ of six equally spaced radiance bins.
The northern region represents disk and limb (top), the southern region
the prominence core (bottom).
}
\label{ProfJun15}
\end{figure}
\begin{figure}
\includegraphics[width=8cm]{f4_13875.eps}
\caption{Idem for the inner (top) and outer (bottom) region in Fig.\,\ref{Jun9}.}
\label{ProfJun9}
\end{figure}
The central reversals of the Ly$-\alpha$~ profiles in both prominences differ, the
profiles obtained on Jun 15 were more reversed than those from Jun 9. This
may be related to the different orientation of the prominence axes as derived
from EIT 304 (cf., Figs.\, 2 and 3) and Kanzelh\"ohe H-$\alpha$ images. On Jun 15,
the threads were rather perpendicular to the line-of-sight, while
more edge-on (LOS parallel to the field lines) on Jun 9. Such an
explanation would be consistent with the model calculations and predictions
of \citet{Heinzel05}. Observational evidence for such a scenario based on spectra
of the higher Lyman lines, Ly2 to Ly7, has been reported by \citet{Schmieder07}.
\section{Radiance histograms}
We already noted the low contrast of the prominences in Ly$-\alpha$. In Fig.\,\ref{HistJun15}
we show radiance histograms of the prominence core and of the on-disk
locations in Fig.\,\ref{Jun15}. For comparison we add as dotted line the log-normal
radiance distribution of Ly$-\alpha$~ in the quiet Sun as presented in earlier work \citep{Curdt08}.
Note that the disk histogram, having a different bin size, was scaled for
better comparison.
Although the small number of prominence pixels only allows a noisy distribution,
the differences are, as expected, very obvious. The histogram is per definition clipped at the
dim side because of the area selection criterion. The main difference is certainly
found in the high-radiance part; the prominence histogram completely lacks
brighter pixels, which makes it a very narrow distribution. The uniform emergent emission translates,
according to the Barbier-Eddington relation, to a uniform source function at
an optical depth, $\tau$, of unity and is indicative of homogeneous populations
of the 1s and 2p levels, and thus rather homogeneous thermodynamic conditions.
In Fig.\,\ref{HistJun9} we show the radiance distribution of the prominence in
Fig.\,\ref{Jun9} in Ly$-\alpha$~ and in Si\,{\sc iii} emission. This data set has more
prominence pixels and allows to also include fainter pixels here.
We defined an empirically determined discrimination level to separate prominence emission
from coronal background and defined the lower-15\,\% radiance category as coronal background, which does not belong
to the prominence. The Ly$-\alpha$~ histogram has a sharp upper limit and, in
contrast to the disk histogram, a low-radiance wing. The Si\,{\sc iii}
histogram of this prominence differs significantly, as one could expect, from
both the quiet Sun state (dotted line) and from its Ly$-\alpha$~ counter part.
We conclude, that the radiance distributions of both prominences are, as a
consequence of dissimilar physical conditions, remarkably different from the
log-normal distribution of the average quiet Sun \citep{Fontenla88,Curdt08}.
\begin{figure}
\includegraphics[width=8cm]{f6_13875.eps}
\caption{Radiance distribution of the limb location (1+2) in Fig.\,\ref{Jun15}
(left) and of the prominence core (right). The low contrast of the prominence location translates
into a narrow distribution, which differs significantly from the log-normal
distribution, which \citet{Curdt08} found in the quiet Sun at disk center (dotted curve).}
\label{HistJun15}
\end{figure}
\begin{figure}
\includegraphics[width=8cm]{f7_13875.eps}
\caption{Idem for the prominence in Fig.\,\ref{Jun9}.
The coronal background of pixels in the lower-15\,\% radiance category has been excluded.}
\label{HistJun9}
\end{figure}
\section{Summary and conclusion}
We have presented the first SUMER observations of prominences in the light of
the hydrogen Ly$-\alpha$~ line at 1216~{\AA} with reduced incoming photon flux to avoid
saturation effects of the SUMER detection system. We have completed a statistical
analysis and report salient empirical results derived thereof.
As such, we found clear evidence in support of models, which predict an effect
of the orientation of the magnetic field relative to the line of sight on
the asymmetry of the Ly$-\alpha$~ profile.
The Lyman lines are more reversed if the line of sight is across the prominence
axis as compared to the case when it is aligned along its axis.
Given the great variability in the appearance of prominences
and the wide range of physical parameters, the observation of two
prominences is by far not enough to cover all the issues. We felt, however, that
our results constitute a piece of information important enough to be presented
here. More joint observations of prominences and modeling of their Ly$-\alpha$~ line profile are
highly desirable.
\begin{acknowledgements}
The SUMER project is financially supported by DLR, CNES, NASA, and the ESA PRODEX
Programme (Swiss contribution). SUMER is part of {\it SOHO} of ESA and NASA.
HT is supported by the International Max Planck Research School for his stay at MPS.
This non-routine observation was performed with the help of D. Germerott. This paper
greatly benefited from very constructive comments of the referee.
\end{acknowledgements}
|
\section{Introduction}
The great challenge of string phenomenology
is constructing realistic string models which allow us to make unique
predictions that can be tested at the Large Hadron Collider (LHC),
future International Linear Collider (ILC), and other
experiments. If these predictions are confirmed at
future experiments, we will possess
strong evidence to support that string theory is indeed the correct
fundamental description of nature.
To the present, string phenomenology has been primarily
concentrated on heterotic $E_8\times E_8$ string theory and
Type II string theories with D-branes, though unfortunately,
this has not resulted in any unique
prediction thus far.
The last two years have seen Grand Unified Theories (GUTs)
constructed locally in F-theory, which can be considered as the
strongly coupled formulation of ten-dimensional Type IIB string
theory with a varying axion-dilaton field
$S$~\cite{Vafa:1996xn, Donagi:2008ca, Beasley:2008dc,
Beasley:2008kw, Donagi:2008kj}. In F-theory model building,
the gauge fields reside on the observable seven-branes that wrap a del Pezzo
$n$ ($dP_n$) surface for the extra four space dimensions,
while the Standard Model (SM) fermions
and Higgs fields are localized on the complex
codimension one curves (two-dimensional real subspaces) in $dP_n$.
Certainly, F-Theory model building and phenomenology have been
studied extensively~\cite{Font:2008id, Jiang:2009zza,
Blumenhagen:2008aw, Heckman:2009bi, Donagi:2009ra,
Jiang:2009za, Li:2009cy, Li:2009fq}.
In contrast to D-brane model building~\cite{Berenstein:2006aj},
all the SM fermion Yukawa couplings
can be obtained from the triple
intersections of the SM fermion and Higgs curves.
An exciting new feature is that $SU(5)$ gauge symmetry
can be broken down to the SM gauge symmetry
by turning on $U(1)_Y$
flux~\cite{Beasley:2008dc, Beasley:2008kw, Li:2009cy},
and additionally, the $SO(10)$ gauge
symmetry can be broken down to the $SU(5)\times U(1)_X$
and $SU(3)\times SU(2)_L\times SU(2)_R\times U(1)_{B-L}$
gauge symmetries by turning on the $U(1)_X$ and $U(1)_{B-L}$
fluxes, respectively~\cite{Beasley:2008dc, Beasley:2008kw,
Jiang:2009zza, Jiang:2009za, Font:2008id, Li:2009cy}.
It is significant to note that realistic GUTs from F-theory can be constructed
locally, hence, the next key question is whether a unique
prediction can be made that can tested at
the LHC, ILC, and other future experiments.
To study the low energy phenomenology from F-theory GUTs,
gauge mediated supersymmetry breaking
was predominantly considered since the F-theory GUTs were constructed
locally~\cite{Heckman:2009bi}. However, to construct realistic F-theory GUTs,
we must embed these local GUTs into a globally consistent
framework~\cite{Donagi:2009ra}.
Consequently, here we study gravity mediated supersymmetry breaking. In F-theory
$SU(5)$ and $SO(10)$ models where the gauge symmetries are
broken down
to the $SU(3)\times SU(2)_L\times U(1)_Y$ and
$SU(3)\times SU(2)_L\times SU(2)_R\times U(1)_{B-L}$
gauge symmetries by turning on the $U(1)_Y$ and $U(1)_{B-L}$
fluxes, respectively, we obtain the exact
gaugino mass relation
(See Eq.~(\ref{GMRelation}) in the following)
near the electroweak scale at one loop.
These F-theory GUTs are constructed locally,
so we do not know the K\"ahler potential for the SM fermions
and Higgs fields. For this reason, we cannot calculate the
supersymmetry breaking
scalar masses and trilinear soft terms.
Interestingly, our gaugino mass relation can be preserved
very well at the low energy two-loop level if the
scalar masses and trilinear soft terms
are near the TeV scale. Using the indices for the gaugino mass
relations~\cite{Li:2010xr, Li:2010hi}, we show that
our gaugino mass relation is different from those that have been
studied thus far~\cite{Choi:2007ka}. Because the gaugino masses can be measured
at LHC and ILC~\cite{Cho:2007qv, Barger:1999tn, Altunkaynak:2009tg},
these F-theory GUTs can be tested at the colliders.
Note that the generic scalar masses and trilinear soft terms will not affect our
prediction on the gaugino mass relation at low energy,
so we assume a universal scalar mass $m_0$ and universal trilinear soft
term $A_{0}$ for simplicity. Examining two typical scenarios of
gaugino masses, we present the viable parameter space which
satisfies all the latest experimental constraints and is consistent with
the CDMS II experiment~\cite{Ahmed:2009zw}. In particular, the gaugino mass
relation is in fact satisfied at two-loop level with only a very slight
deviation at low energy. More detailed discussions will be
presented elsewhere~\cite{LMN-Preparation}.
\section{Gaugino Mass Relation}
In the F-theory GUTs, the GUT gauge symmetries on
the observable seven-branes are broken by turning on
the $U(1)$ fluxes. Interestingly, these $U(1)$ fluxes will
give additional
contribution to the gauge kinetic functions, which can be computed
by dimensionally reducing the Chern-Simons action of the
observable seven-branes wrapping on $dP_n$
\begin{eqnarray}
S_{\rm CS} &=& \mu_7 \int_{dP_n\times \mathbb{R}^{3,1}} a \wedge {\rm tr}(F^4) ~.~\,
\end{eqnarray}
For simplicity, we will assume that the heavy
KK states and string states have masses above the GUT scale,
which can be realize naturally in the global F-theory GUTs.
First, let us consider the $SU(5)$
models~\cite{Beasley:2008dc, Beasley:2008kw, Li:2009cy}.
Turning on the $U(1)_Y$ flux, the gauge kinetic functions
$f_3$, $f_2$ and $f_1$ respectively
for $SU(3)_C$, $SU(2)_L$ and $U(1)_Y$ gauge symmetries
at the string scale can be
parametrized as follows~\cite{Donagi:2008kj, Blumenhagen:2008aw}
\begin{eqnarray}
f_3 &=& \tau + {1 \over 2} \alpha S~,~\,
f_2 ~=~ \tau + {1 \over 2} \left(\alpha+2 \right) S~,~ \nonumber \\
f_1 &=& \tau + {1 \over 2} \left(\alpha+{6\over 5} \right) S~,~\,
\end{eqnarray}
where $\tau$ is the original gauge kinetic function of
$SU(5)$, the $S$ terms arise from $U(1)_Y$ flux contributions,
and $\alpha$ is a positive integer.
Second, let us consider the $SO(10)$ models. If the $SO(10)$ gauge symmetry
is broken down to the flipped $SU(5)\times U(1)_X$ gauge symmetry
by turning on the $U(1)_X$ flux~\cite{Beasley:2008dc, Beasley:2008kw,
Jiang:2009zza, Jiang:2009za}, we can show that the gauge kinetic
functions for $SU(5)$ and $U(1)_X$ are exactly the
same at the unification scale~\cite{Jiang:2009za}. Interestingly,
if we break the $SO(10)$ gauge symmetry down to the
$SU(3)\times SU(2)_L\times SU(2)_R\times U(1)_{B-L}$
gauge symmetry by turning on the $U(1)_{B-L}$
flux~\cite{Font:2008id, Li:2009cy},
we can show that the gauge kinetic functions
for the $SU(3)_C$, $U(1)_{B-L}$, $SU(2)_L$, and $SU(2)_R$
gauge symmetries at the string scale are~\cite{Li:2009cy}
\begin{eqnarray}
f_{SU(3)_C} &=& f_{U(1)_{B-L}}= \tau + S ~,~\, \nonumber \\
f_{SU(2)_L} &=& f_{SU(2)_R} = \tau ~,~\,
\end{eqnarray}
where $\tau$ is the original gauge kinetic function of $SO(10)$,
and the $S$ term arises from $U(1)_{B-L}$ flux contribution.
We can break the $SU(2)_R\times U(1)_{B-L}$ gauge symmetry
down to $U(1)_Y$ at the string scale by the Higgs mechanism.
As a consequence, we obtain the gauge kinetic function
for $U(1)_Y$~\cite{Li:2009cy}
\begin{eqnarray}
f_{U(1)_{Y}}= {3\over 5} f_{SU(2)_R} + {2\over 5} f_{U(1)_{B-L}}=
\tau + {2\over 5} S ~.~\,
\end{eqnarray}
Now, let us study gravity mediated supersymmetry breaking.
We can show that the gaugino mass relation in
the $SO(10)$ models with $U(1)_{B-L}$ flux
is the same as that in the $SU(5)$ models with $U(1)_Y$ flux.
Henceforth, we only consider
the $SU(5)$ models with $U(1)_Y$ flux here.
Supposing supersymmetry is broken by the F-terms of $\tau$ and $S$,
we can parametrize $F^\tau$ and $F^S$ as follows
\begin{eqnarray}
F^{\tau} = M_{3/2}^{\prime} (\tau+\overline{\tau}) \cos\theta ~,~\,
F^S = M_{3/2}^{\prime} (S+\overline{S}) \sin\theta~,~\,
\end{eqnarray}
where $ M_{3/2}^{\prime}$ is the gravitino mass if supersymmetry
is only broken by the F-terms of $\tau$ and $S$.
Then, the gaugino masses $M_3$, $M_2$, and $M_1$ respectively
for $SU(3)_C$, $SU(2)_L$, and $U(1)_Y$
at the GUT scale are
\begin{eqnarray}
M_3 &=& {{ \cos\theta + \alpha x \sin\theta} \over\displaystyle
{1 + \alpha x}} M^{\prime}_{3/2} ~,~\, \nonumber \\
M_2 &=& {{ \cos\theta + (\alpha + 2)x \sin\theta} \over\displaystyle
{1+(\alpha+2)x}} M^{\prime}_{3/2} ~,~\, \nonumber \\
M_1 &=& {{5 \cos\theta + (5\alpha + 6)x \sin\theta} \over\displaystyle
{5+(5\alpha+6)x}} M^{\prime}_{3/2} ~,~\,
\end{eqnarray}
where $x$ is defined as
\begin{eqnarray}
x &=& {{S+\overline{S}} \over {2(\tau +\overline{\tau})}} ~.~\,
\end{eqnarray}
Using the one-loop renormalization group equations (RGEs), we obtain
the gaugino mass relation around the electroweak scale
\begin{eqnarray}
{{M_1}\over {\alpha_1}} - {{M_3}\over {\alpha_3}}
&=& {3\over 5} \left({{M_2}\over {\alpha_2}} - {{M_3}\over {\alpha_3}} \right) ~,~\,
\label{GMRelation}
\end{eqnarray}
where $\alpha_3$, $\alpha_2$, and $\alpha_1$ are the gauge couplings for
the $SU(3)_C$, $SU(2)_L$, and $U(1)_Y$ gauge symmetries, respectively.
Following Refs.~\cite{Li:2010xr, Li:2010hi}, we define the index $k$
for the gaugino mass relation
as follows
\begin{eqnarray}
k &=& {{{M_2} {\alpha^{-1}_2} - {{M_3} {\alpha^{-1}_3}}}\over
{{{M_1} {\alpha^{-1}_1}} - {{M_3} {\alpha^{-1}_3}}}} ~.~
\end{eqnarray}
Thus, in the F-theory $SU(5)$ and $SO(10)$ models respectively
with $U(1)_Y$ and $U(1)_{B-L}$ fluxes, we obtain that
the index for gaugino mass relation is $5/3$, {\it i.e.}, $k=5/3$.
Moreover, in the minimal supersymmetric Standard Model with
anomaly mediation and mirage mediation~\cite{Choi:2007ka}, we can show that
the index for gaugino mass relation is $5/12$, {\it i.e.}, $k=5/12$~\cite{Li:2010hi}.
Thus, we emphasize that our gaugino mass relation is
different from those in simple anomaly mediation and
mirage mediation~\cite{Choi:2007ka}.
Furthermore, the index for gaugino mass relation in the minimal
Supergravity (mSUGRA)~\cite{Choi:2007ka} is not well defined
but can be formally written
as $0/0$, {\it i.e.}, $k=0/0$~\cite{Li:2010xr}.
So the gaugino mass relation in mSUGRA
satisfies the above gaugino mass relation in
Eq. (\ref{GMRelation}). However,
if $2(M_1\alpha_1^{-1}-M_3 \alpha_3^{-1})/(M_1\alpha_1^{-1}+M_3 \alpha_3^{-1})$
is not very small,
our gaugino mass relation can definitely be distinguished from that
of mSUGRA. Moreover, the gaugino masses can be measured at
the LHC and ILC~\cite{Cho:2007qv, Barger:1999tn, Altunkaynak:2009tg}. Therefore,
these F-theory GUTs can be tested at the LHC and ILC, and may be
distinguished from the mSUGRA, simple anomaly mediation and mirage
mediation~\cite{Choi:2007ka}.
To test the gaugino mass relation close to the electroweak scale, we define a
parameter $\eta$ as follows
\begin{eqnarray}
\eta &=& {{ 5\left({{M_1} {\alpha^{-1}_1}} - {{M_3} {\alpha^{-1}_3}}\right)}
\over
{ 3\left({{M_2} {\alpha^{-1}_2}} - {{M_3} {\alpha^{-1}_3}} \right)}} ~.~\,
\label{Eq-eta}
\end{eqnarray}
Notice $\eta$ is exactly one at the GUT scale. In addition, $\eta$ is one
around the electroweak scale from one-loop RGE running, yet $\eta$ may deviate slightly
from one as a result of two-loop RGE running.
For simplicity, we assume that $x$ is small in this work, and then
we have approximate gauge coupling unification
at the GUT scale, allowing us to use well-established public codes
for computations. For gaugino masses, we consider two typical
scenarios
(I) We consider the dilaton dominated scenario, {\it i.e.},
$\theta=\pi/2$. The gaugino
masses at the GUT scale are
\begin{eqnarray}
M_3 & \simeq & \alpha M_{1/2} ~,~\,
M_2 ~ \simeq ~ \left( \alpha + 2 \right) M_{1/2} ~,~\, \nonumber \\
M_1 & \simeq & \left( \alpha + {6\over 5} \right) M_{1/2} ~,~\,
\end{eqnarray}
where $M_{1/2}$ is a mass parameter. In our numerical calculations,
we will choose $\alpha=3$.
(II) We consider the scenario where
$\cos\theta$ is on the order of $ x \sin\theta$.
Assuming $\cos\theta > 0$ and $\sin\theta < 0$,
we parametrize
$\cos\theta$ as follows
\begin{eqnarray}
\cos\theta & = & - \gamma x \sin\theta ~,~\,
\end{eqnarray}
where $\gamma $ is a positive real number.
Thus, we obtain the gaugino
masses at the GUT scale
\begin{eqnarray}
M_3 & \simeq & \left( \gamma - \alpha \right) M_{1/2} ~,~\,
M_2 ~ \simeq ~ \left( \gamma - \alpha - 2 \right) M_{1/2} ~,~\, \nonumber \\
M_1 & \simeq & \left( \gamma - \alpha - {6\over 5} \right) M_{1/2} ~.~\,
\end{eqnarray}
In our numerical calculations,
we choose $(\gamma - \alpha)= 5$. In summary,
we have $M_3 < M_1 < M_2 $ in scenario I and $M_2 < M_1 < M_3$ in scenario II.
\section{Low Energy Supersymmetry Phenomenology}
We take $\mu > 0$,
so we have four free parameters in our models:
$M_{1/2}$, $m_0$, $A_{0}$, and $\tan\beta$, where
$\tan\beta$ is the ratio of the Higgs vacuum expectation values.
The soft supersymmetry breaking terms are input into
{\tt MicrOMEGAs 2.0.7}~\cite{Belanger:2006is}
using {\tt SuSpect 2.34}~\cite{Djouadi:2002ze} as a front end to
run the soft terms down to the electroweak scale via RGEs
and then to calculate the corresponding neutralino relic
density. We use a top quark mass
of $m_{t}$ = 173.1 GeV~\cite{:2009ec}. The direct detection cross-sections are
calculated using {\tt MicrOMEGAs 2.1}~\cite{Belanger:2008sj}. We employ the
following experimental constraints: (1) The WMAP $2\sigma$ measurements of
the cold dark matter density~\cite{Spergel:2006hy},
0.095 $\leq \Omega_{\chi} \leq$ 0.129. Also, we allow $\Omega_{\chi}$ to be
larger than the upper bound due to a possible $\cal{O}$(10) dilution factor~\cite{Mavromatos:2009pm}
and to be smaller than the lower bound due to multicomponent
dark matter. (2) The experimental limits on
the Flavor Changing Neutral Current (FCNC) process, $b \rightarrow s\gamma$.
The results from the Heavy Flavor Averaging Group (HFAG)~\cite{Barberio:2007cr},
in addition to the BABAR, Belle, and CLEO results, are:
$Br(b \rightarrow s\gamma) = (355 \pm 24^{+9}_{-10} \pm 3) \times 10^{-6}$.
There is also a more recent estimate~\cite{Misiak:2006zs}
of $Br(b \rightarrow s\gamma) = (3.15 \pm 0.23) \times 10^{-4}$. For our analysis,
we use the limits
$2.86 \times 10^{-4} \leq Br(b \rightarrow s\gamma) \leq 4.18 \times 10^{-4}$,
where experimental and theoretical errors are added in quadrature.
(3) The anomalous magnetic moment of the muon, $g_{\mu} - 2$.
For this analysis we use the $2\sigma$ level boundaries,
$11 \times 10^{-10} < \Delta a_{\mu} < 44 \times 10^{-10}$~\cite{Bennett:2004pv}.
(4) The process $B_{s}^{0} \rightarrow \mu^+ \mu^-$ where the decay
has a $\mbox{tan}^6\beta$ dependence. We take the upper bound to be
$Br(B_{s}^{0} \rightarrow \mu^{+}\mu^{-}) < 5.8 \times 10^{-8}$~\cite{:2007kv}.
(5) The LEP limit on the lightest CP-even Higgs boson
mass, $m_{h} \geq 114$ GeV~\cite{Barate:2003sz}.
For scenario I, we commence with $m_{0}$, $A_{0}$, and tan$\beta$ as free parameters,
however, a comprehensive scan uncovers $A_{0} = m_{0}$ as the most phenomenologically
favored. As shown in Fig.~\ref{fig:mo_scenario1}, the experimentally allowed parameter
space for the Scenario I with $\alpha$ = 3, $\beta$ = 2, and tan$\beta$ = 51 after
applying all these constraints consists of small $M_{1/2}$ and large $m_{0}$.
We choose a point within the narrow region that satisfies the WMAP relic density
as our benchmark point for analysis. See table~\ref{tab:Scenario1_masses} for the
supersymmetric particle (sparticle) and Higgs spectrum. In fact, with a relic density
of $\Omega_{\chi}$ = 0.1156,
this benchmark point additionally satisfies the very constrained WMAP 5-year
results~\cite{Spergel:2006hy}. For constant $m_{0} = A_{0}$ = 740 GeV,
we find tan$\beta$ = 25-52 for $0 \leq \Omega_{\chi} \lesssim 1.1$, in contrast
to tan$\beta$ = 41 and tan$\beta$ = 51-52 for the WMAP region.
In mSUGRA, the focus point region consists of large $m_{0}$ where the WMAP observed
relic density can be satisfied with a large Higgsino component in the lightest
supersymmetric particle (LSP) neutralino due to a small $|\mu|$. However, even
though $m_{0}$ is reasonably large in comparison to $M_{1/2}$ for this benchmark
point, here the LSP is 98\% bino. The WIMP-nucleon direct-detection
cross-sections $\sigma_{SI}$ depicted in Fig.~\ref{fig:cdms_scenario1} underscore the fact
that the case of $\alpha$ = 3, $\beta$ = 2, and tan$\beta$ = 51 produces WIMPs
with $\sigma_{SI}$ just beneath the CDMS II~\cite{Ahmed:2009zw} upper limit, with
our benchmark point at $\sigma_{SI} = 6.15 \times 10^{-8}$ pb and
$m_{\widetilde{\chi}_{1}^{0}}$ = 316 GeV. The constraints from previous
ZEPLIN~\cite{Lebedenko:2008gb}, XENON~\cite{Angle:2007uj}, and CDMS~\cite{Ahmed:2008eu}
experiments are also delineated on the plot.
\begin{table}[ht]
\small
\centering
\caption{Supserysmmetric particle (sparticle)
and Higgs spectrum for a Scenario I, $\alpha = 3,~\beta = 2$ benchmark
point with $\sigma_{SI} = 6.15 \times 10^{-8}$ pb. Here, tan$\beta$ = 51 and
$\Omega_{\chi}$ = 0.1156. The GUT scale mass parameters for this point are (in Gev)
$M_{1/2}$ = 180, $M_3$ = 540, $M_2$ = 900, $M_1$ = 756, $m_{0} = A_{0}$ = 740.}
\begin{tabular}{|c|c||c|c||c|c||c|c||c|c||c|c|} \hline
$\widetilde{\chi}_{1}^{0}$&$316$&$\widetilde{\chi}_{1}^{\pm}$&$473$&$\widetilde{e}_{R}$&$790$&$\widetilde{t}_{1}$&$973$&$\widetilde{u}_{R}$&$1302$&$m_{h}$&$115.4$\\ \hline
$\widetilde{\chi}_{2}^{0}$&$477$&$\widetilde{\chi}_{2}^{\pm}$&$743$&$\widetilde{e}_{L}$&$946$&$\widetilde{t}_{2}$&$1201$&$\widetilde{u}_{L}$&$1402$&$m_{A}$&$465$\\ \hline
$\widetilde{\chi}_{3}^{0}$&$487$&$\widetilde{\nu}_{e/\mu}$&$942$&$\widetilde{\tau}_{1}$&$489$&$\widetilde{b}_{1}$&$1103$&$\widetilde{d}_{R}$&$1294$&$m_{H^{\pm}}$&$473$\\ \hline
$\widetilde{\chi}_{4}^{0}$&$743$&$\widetilde{\nu}_{\tau}$&$837$&$\widetilde{\tau}_{2}$&$843$&$\widetilde{b}_{2}$&$1195$&$\widetilde{d}_{L}$&$1404$&$\widetilde{g}$&$1263$\\ \hline
\end{tabular}
\label{tab:Scenario1_masses}
\end{table}
\begin{figure}[ht]
\centering
\includegraphics[width=0.4\textwidth]{mo_Scenario1.eps}
\caption{Experimentally allowed parameter space for Scenario I,
$\alpha = 3,~\beta = 2$, $A_{0} = m_{0}$, tan$\beta$ = 51. The benchmark point detailed
in Table~\ref{tab:Scenario1_masses} is annotated on the plot by the orange point.}
\label{fig:mo_scenario1}
\end{figure}
\begin{figure}[ht]
\centering
\includegraphics[width=0.4\textwidth]{cdms_scenario1.eps}
\caption{Spin-independent WIMP-nucleon cross-sections for Scenario I,
$\alpha = 3,~\beta = 2$, $A_{0} = m_{0}$, tan$\beta$ = 51. The green shaded region
satisfies all experimental constraints. The point detailed
in Table~\ref{tab:Scenario1_masses} is annotated on the plot by the orange point.}
\label{fig:cdms_scenario1}
\end{figure}
The values of $M_1, ~M_2, ~M_3, ~\alpha_{1}, ~\alpha_{2}$ and $\alpha_{3}$ at
electroweak-symmetry breaking (EWSB) are used to compute $\eta$ in Eq.~(\ref{Eq-eta}),
and we find the deviation of $\eta$ from one is very small, or on the order
of 1.2\% - 1.6\%, as expected. The small deviation from one for $\alpha$ = 3,
$\beta$ = 2, and tan$\beta$ = 51 is clearly shown in Fig.~\ref{fig:eta_scenario1},
thus, the gaugino mass relation in Eq. (\ref{GMRelation}) can be tested
at the LHC and ILC since the two-loop corrections are indeed very small.
\begin{figure}[ht]
\centering
\includegraphics[width=0.4\textwidth]{eta1.eps}
\caption{Plot of $\eta$ for Scenario I,
$\alpha = 3,~\beta = 2$, $A_{0} = m_{0}$, tan$\beta$ = 51. The shaded regions satisfy
all experimental constraints. The point detailed in Table~\ref{tab:Scenario1_masses}
is annotated on the plot by the orange point.}
\label{fig:eta_scenario1}
\end{figure}
The five lightest sparticles for the $\alpha$ = 3, $\beta$ = 2, and tan$\beta$ = 51
benchmark point, including the heavy Higgs, are
$\widetilde{\chi}_{1}^{0} < A < H^{\pm} < \widetilde{\chi}_{1}^{\pm} < \widetilde{\chi}_{2}^{0}$.
The productions of squarks $\widetilde{q}$ and gluino $\widetilde{g}$ have
the largest differential cross-sections at LHC. The squark in the first two families
will decay dominantly into the gluino $\widetilde{g}$ and the corresponding quark, and
the gluino $\widetilde{g}$ decay will mainly produce either the sbottom $\widetilde{b}$ and bottom quark $b$
or the stop $\widetilde{t}$ and top quark $t$. The $\widetilde{b}$ and
$\widetilde{t}$ decay to the top quark $t$, bottom quark $b$, neutralino, and
chargino. Additionally, we will get $b$ quark from $\widetilde{\chi}_{2}^{0}$ through
light Higgs in the process $\widetilde{\chi}_{2}^{0} \rightarrow h_{0} \widetilde{\chi}_{1}^{0}$
with a branching ratio of 93\%. These light Higgs will in turn decay to $b \overline{b}$
with a 73\% branching ratio. Leptons and hadronic jets will result from the
decay $\widetilde{\chi}_{1}^{\pm} \rightarrow W^{\pm} \widetilde{\chi}_{1}^{0}$, where
this is the only kinematically allowed $\widetilde{\chi}_{1}^{\pm}$ process. Due to
the less massive nature of the heavy Higgs particles, the $q \overline{q}$ reaction
will provide some neutral heavy Higgs field $A$ that will decay to $b \overline{b}$
pair 87\% of the time, while the heavy charged Higgs field $H^{\pm}$ will
produce $\overline{b}t$ or $b\overline{t}$ pair 84\% of the time,
with $t \rightarrow W^{+}b$. Thus, this benchmark point will produce mainly
light Higgs field $h_{0}$, $b$ quark, and $W$ boson at LHC.
In addition to the viable parameter space of Scenario I, Scenario II can also generate
a well constrained region at the relic densities we consider here. Exhibited in
Fig.~\ref{fig:mo_scenario2} is the experimentally allowed region of
small $m_{0}$ and $M_{1/2}$ for Scenario I with
$\gamma - \alpha = 5,~\beta = 2$, $A_{0} = m_{0}$,
and tan$\beta$ = 27. Choosing a benchmark point with
$M_{1/2}=110$ GeV and $m_0=A_0=190$ GeV, we present the sparticle and Higgs spectrum
in Table~\ref{tab:Scenario2_masses}. Because the small mass difference of 10 GeV between
the LSP neutralino $\widetilde{\chi}_{1}^{0}$
and the next to the lightest supersymmetric particle (NLSP) $\widetilde{\tau}_{1}^{\pm}$,
the LSP neutralino in the early Universe can annihilate with stau and then we can
obtain the correct dark matter density, similar to
the stau-neutralino coannihilation region in mSUGRA.
Moreover, the LSP for this point is 99.7\% bino.
Considering a constant $m_{0} = A_{0}$ = 190 GeV, we discover tan$\beta$ = 8-59
for $0 \leq \Omega_{\chi} \lesssim 1.1$, but only tan$\beta$ = 18-27 for the WMAP region.
This Scenario II benchmark point possesses a $\sigma_{SI} = 2.03 \times 10^{-8}$ pb
at $m_{\widetilde{\chi}_{1}^{0}}$ = 170 GeV, near the CDMSII upper limit. Furthermore,
a calculation of $\eta$ in Eq.~(\ref{Eq-eta}) for this Scenario II benchmark point yields
comparable results to the Scenario I benchmark point, namely only a very small deviation
from one, on the order of 1.5\% to 3.5\%, as depicted in Fig.~\ref{fig:eta_scenario2},
corroborating the delineation of $\eta$ in Fig.~\ref{fig:eta_scenario1} and
its testability. A close examination of this benchmark point
reveals $\widetilde{\chi}_{2}^{0} \rightarrow \widetilde{\tau}_{1}^{\mp}
\tau^{\pm} \rightarrow \tau^{\mp} \tau^{\pm} \widetilde{\chi}_{1}^{0}$ as the dominant decay,
therefore, we would expect opposite sign tau pair to be characteristic at the LHC.
\begin{table}[ht]
\small
\centering
\caption{Sparticle and Higgs spectrum for a Scenario II, $\gamma - \alpha = 5,~\beta = 2$
benchmark point with $\sigma_{SI} = 2.03 \times 10^{-8}$ pb. Here, tan$\beta$ = 27 and
$\Omega_{\chi}$ = 0.107. The GUT scale mass parameters for this point are (in Gev)
$M_{1/2}$ = 110, $M_3$ = 550, $M_2$ = 330, $M_1$ = 418, $m_{0} = A_{0}$ = 190.}
\begin{tabular}{|c|c||c|c||c|c||c|c||c|c||c|c|} \hline
$\widetilde{\chi}_{1}^{0}$&$170$&$\widetilde{\chi}_{1}^{\pm}$&$256$&$\widetilde{e}_{R}$&$247$&$\widetilde{t}_{1}$&$922$&$\widetilde{u}_{R}$&$1117$&$m_{h}$&$115.7$\\ \hline
$\widetilde{\chi}_{2}^{0}$&$256$&$\widetilde{\chi}_{2}^{\pm}$&$687$&$\widetilde{e}_{L}$&$293$&$\widetilde{t}_{2}$&$1080$&$\widetilde{u}_{L}$&$1126$&$m_{A}$&$661$\\ \hline
$\widetilde{\chi}_{3}^{0}$&$681$&$\widetilde{\nu}_{e/\mu}$&$283$&$\widetilde{\tau}_{1}$&$180$&$\widetilde{b}_{1}$&$1033$&$\widetilde{d}_{R}$&$1115$&$m_{H^{\pm}}$&$666$\\ \hline
$\widetilde{\chi}_{4}^{0}$&$686$&$\widetilde{\nu}_{\tau}$&$276$&$\widetilde{\tau}_{2}$&$321$&$\widetilde{b}_{2}$&$1095$&$\widetilde{d}_{L}$&$1129$&$\widetilde{g}$&$1257$\\ \hline
\end{tabular}
\label{tab:Scenario2_masses}
\end{table}
\begin{figure}[ht]
\centering
\includegraphics[width=0.4\textwidth]{mo_Scenario2.eps}
\caption{Experimentally allowed parameter space for Scenario II,
$\gamma - \alpha = 5,~\beta = 2$, $A_{0} = m_{0}$, tan$\beta$ = 27. The benchmark point
detailed in Table~\ref{tab:Scenario2_masses} is annotated on the plot by the orange point.
Identification of the excluded regions is shown in the chart legend in Fig.~\ref{fig:mo_scenario1}.}
\label{fig:mo_scenario2}
\end{figure}
\begin{figure}[ht]
\centering
\includegraphics[width=0.4\textwidth]{cdms_scenario2.eps}
\caption{Spin-independent WIMP-nucleon cross-sections for Scenario II,
$\gamma - \alpha = 5,~\beta = 2$, $A_{0} = m_{0}$, tan$\beta$ = 27. The green shaded region
satisfies all experimental constraints. The point detailed in Table~\ref{tab:Scenario2_masses}
is annotated on the plot by the orange point.}
\label{fig:cdms_scenario2}
\end{figure}
\begin{figure}[ht]
\centering
\includegraphics[width=0.4\textwidth]{eta2.eps}
\caption{Plot of $\eta$ for Scenario II,
$\gamma - \alpha = 5,~\beta = 2$, $A_{0} = m_{0}$, tan$\beta$ = 27. The shaded regions
satisfy all experimental constraints. The point detailed in Table~\ref{tab:Scenario2_masses} is
annotated on the plot by the orange point.}
\label{fig:eta_scenario2}
\end{figure}
Let us comment on the phenomenological differences between our models and the other
supersymmetry breaking mediation models. Because
our gaugino mass relation is different from those in the
mSUGRA, simple anomaly mediation and mirage mediation, our gaugino mass ratio
$M_1 : M_2 : M_3$ at the low energy are different from those in the
mSUGRA, simple anomaly mediation and mirage mediation. And then the neutralino masses,
chargino masses and gluino mass will be different as well, which will affect
the dark matter density and the productions and decays of the
supersymmetric particles at the LHC.
\section{Conclusion}
We considered gravity mediated
supersymmetry breaking and
derived the exact gaugino mass relation at one loop
near the electroweak scale in the F-theory
$SU(5)$ and $SO(10)$ models with $U(1)_Y$ and $U(1)_{B-L}$
fluxes, respectively. The gaugino mass relation presented in this work
differs from the typical gaugino mass relations that
have been studied in the past, and should be preserved
pretty well at low energy in general. Thus,
these F-theory GUTs can be tested at the LHC
and forthcoming ILC. We exhibited two concrete scenarios
that satisfy all the latest
experimental constraints and are consistent with the
CDMS II experiment. Most importantly, the gaugino
mass relation is indeed satisfied at two-loop level
with only a very small deviation.
\section*{Acknowledgments}
\begin{acknowledgments}
This research was supported in part
by the DOE grant DE-FG03-95-Er-40917 (TL and DVN),
by the Natural Science Foundation of China
under grant numbers 10821504 and 11075194 (TL),
and by the Mitchell-Heep Chair in High Energy Physics (JM).
\end{acknowledgments}
|
\section[]{Introduction}\label{Intro}
Galaxies reside in three different environments: clusters, groups and voids. The
majority, more than 50 per cent, of galaxies are found in groups and clusters
\citep{Humason:1956p593,Huchra:1982p1,Geller:1983p576,Nolthenius:1987p357,Ramella:2002p514}.
This is particularly true for elliptical galaxies, as the morphology-density
relation has shown \citep{Oemler:1974p646, Dressler:1980p645}. The merger
hypothesis \citep{Toomre:1972p619} suggests that the product of the merger of two
spiral galaxies will be an elliptical galaxy. If this holds, then the probability
to find an elliptical galaxy grows in environments with high densities and low
velocity dispersions, e.g. groups of galaxies. Therefore, local environment is
thought to play a crucial role in galaxy formation and evolution
\citep[e.g.][]{Farouki:1981p656, Balogh:1999p657, Kauffmann:1999p658,
Lemson:1999p654, Moore:1999p653}.
In observations it is relatively easy to study cluster and group galaxies.
Because of this and the potential formation mechanism of elliptical galaxies, it
is not surprising that most studied ellipticals are found in clusters and groups.
Therefore, the detailed properties of isolated field elliptical (IfE) galaxies
have not been extensively studied or very well understood; there are only a few
observational studies and the surveys are small. Moreover, the formation
mechanisms and evolutionary paths of these ``lonely'' elliptical galaxies are not
yet well understood.
In the past two decades several observational projects have identified and
studied the properties of isolated field elliptical galaxies \citep[e.g.][and
references therein]{Smith:2004p250, Reda:2004p502, HernandezToledo:2008p648,
Norberg:2008p652}. In many studies, based on observational data, different
possible formation scenarios have been proposed \citep[see
e.g.][]{Mulchaey:1999p563, Reda:2004p502, Reda:2007p559}, ranging from clumpy
collapse at an early epoch to multiple merging events. Also equal-mass mergers of
two massive galaxies or collapsed groups have been suggested. Theoretical studies
predict that isolated ellipticals are formed in relatively recent mergers of
spiral galaxy pairs, while large isolated ellipticals may be the result of
merging of a small group of galaxies \citep[e.g.][]{Jones:2000p624,
DOnghia:2005p128}.
Observational studies have shown that several IfEs reveal a number of features
such as tidal tails, dust, shells, discy and boxy isophotes and rapidly rotating
discs \citep[e.g.][]{Reduzzi:1996p622, Reda:2004p502, Reda:2005p561,
Hau:2006p557, HernandezToledo:2008p648} indicating recent merger and/or accretion
events. \citet{HernandezToledo:2008p648} concluded that at least $78$ per cent of
their isolated elliptical galaxies show some kind of mophological distortion, and
suggested that these galaxies suffered late dry mergers.
\citet{Reda:2004p502, Reda:2005p561} compiled a sample of $36$ candidates of
isolated early-type galaxies and studied their properties, and concluded that a
major merger of two massive galaxies could explain most observed features. They
also concluded that a collapsed poor group of a few galaxies is a possible
formation scenario. However, \citet{Marcum:2004p572} studied a similar sample of
isolated early-type galaxies, and concluded that isolated systems are
underluminous by at least a magnitude compared with objects identified as merged
group remnants. \citet{Reda:2007p559} concluded that mergers at different
redshifts of progenitors of different mass ratios and gas fractions are needed to
reproduce the observed properties of IfEs.
However, some IfEs have not shown any evidence of recent merging activity
\citep[e.g.][]{Aars:2001p564, Denicolo:2005p568}. \cite{Aars:2001p564}, who
studied a sample of nine isolated elliptical galaxies, identified five galaxies
that were located in environments similar to those of loose groups, while the
environments of the remaining four galaxies were confirmed to be on low density.
All galaxies showed smooth, azimuthally symmetric profiles, with no obvious
indications of dust lanes, nascent spiral structure or star-forming regions. It
is possible that the merging events have happened in distant past, and all signs
of merging events have been wiped out. \cite{Mihos:1995p596} found out, by using
a combination of numerical simulation and synthesized Hubble Space Telescope
(HST) Wide-Field Planetary Camera 2 (WFPC2) images, that merger remnants appear
morphologically indistinguishable from a ``typical'' elliptical $\leq 1$ Gyrs
after the galaxies merged, while \cite{Combes:1995p663} estimated from numerical
simulations that the time might be even less than $0.5$ Gyrs. Indeed, these times
are very short in time scales of galaxy evolution, making it difficult to find
observational evidence to back up formation via merging events. On the other
hand, it is equally possible that some or even all of these galaxies have
initially formed in underdense regions and developed quietly without any major
mergers.
Observational studies of isolated field elliptical galaxies use similar isolation
criteria as optical studies of fossil groups. Both classes show a large magnitude
gap between the first- and the second-ranked galaxy. However, there is no
criterion for IfEs that would require a presence of extended X-ray emission, as
in case of fossil groups. Additionally, fossil groups are not necessarily found
in low-density environments \citep{vonBendaBeckmann:2008p598} like IfEs. Despite
the differences, it is possible that the formation mechanisms of the two systems
are similar or related to each other. Therefore, it is interesting to compare
isolated field elliptical galaxies and fossil groups, and see if these systems
share a common origin.
In this paper, we use the Millennium Simulation \citep{Springel:2005p595}
together with a semi-analytical model \citep{DeLucia:2007p414} of galaxy
formation within dark matter haloes to identify isolated field elliptical
galaxies, to study their properties and formation history, and to compare them
with observational data. This paper is organised as follows. In Section
\ref{Sample}, we discuss the Millennium Simulation, the semi-analytical galaxy
catalogue used, and sample selection. In Section \ref{Results}, we compare the
properties of the Millennium IfEs with observations, while in Section
\ref{ResultsA} we show that the IfEs form a population that is different from the
regular ellipticals. In Section \ref{Formation} we concentrate on evolution of
IfEs and possible formation mechanisms, while in Section \ref{s:discussion} we
discuss our results, the formation of field elliptical galaxies and their
possible connection to fossil groups. Finally, in Section \ref{summary} we
summarise our results and draw the conclusions. Throughout this paper, we adopt a
parametrized Hubble constant: $H_{0} = 100 h$ km s$^{-1}$ Mpc$^{-1}$.
\section[]{Sample Selection}\label{Sample}
\subsection[]{The Millennium Simulation}\label{MS}
We use a simulation that covers a big enough spatial volume and has a
sufficient mass resolution -- the dark matter only Millennium Simulation
\citep[MS;][]{Springel:2005p595}, a $2160^{3}$-particle model of a co-moving cube
of size $500 h^{-1}$ Mpc, on top of which a publicly available semi-analytical
galaxy model \citep{DeLucia:2007p414} has been constructed. The cosmological
parameters used in the MS were: $\Omega_{m} = \Omega_{dm} + \Omega_{b} = 0.25$,
$\Omega_{b} = 0.045$, $h = 0.73$, $\Omega_{\Lambda} = 0.75$, $n = 1$, and
$\sigma_{8} = 0.9$ where the Hubble constant is characterized as $100 h$ km
s$^{-1}$ Mpc$^{-1}$ \citep[for detailed description of the MS,
see][]{Springel:2005p595}. These values were inferred from the first-year WMAP
(Wilkinson Microwave Anisotropy Probe) observations \citep{Spergel:2003p124}.
The galaxy and dark matter halo formation modeling of the MS data is based on
merger trees, built from $64$ individual snapshots. From these time-slices,
merger trees are built by combining the information of all dark matter haloes
found at any given output time. This enables us to trace the formation history
and growth of haloes and subhaloes through time. Properties of galaxies in the MS
data are obtained by using semi-analytic galaxy formation models (SAMs), where
star formation and its regulation by feedback processes is parametrized in
terms of analytical physical models. The semi-analytical galaxy catalogue we use
contains about nine million galaxies at $z = 0$ down to a limiting absolute
magnitude of $M_{R} - 5 \log h = -16.6$. A detailed description of the creation
of the MS Galaxy Catalogue, used in this study, can be found in
\citet{DeLucia:2007p414}.
Semi-analytical galaxy formation models are known to be less than perfect. There
are several free parameters in each SAM that all have a different impact on the
properties of galaxies. The free parameters usually control the feedback effects
and cooling of hot gas. The SAM also controls how gas is stripped from a galaxy
when it approaches another galaxy, and possible starburst events. The physics of
these processes described is not well known. However, in general, SAMs can
reproduce the observed galaxy luminosity function and other statistics well. As
our study is statistical by nature, it is likely that the SAM used for the
Millennium Simulation Galaxy Catalogue is accurate enough to predict the
properties of IfEs and their evolution and formation history.
The MS Galaxy Catalogue does not predict galaxy morphologies. To assign a
morphology for every galaxy, we use their bulge-to-disk ratios.
\citet{Simien:1986p431} found a correlation between the $B-$band bulge-to-disc
ratio and the Hubble type $T$ of galaxies from observations. The mean relation
may be written:
\begin{equation}\label{eq:T}
<\Delta m(T)> = 0.324x(T) - 0.054x(T)^{2} + 0.0047x(T)^{3} ,
\end{equation}
where $\Delta m(T)$ is the difference between the bulge magnitude and the total
magnitude and $x(T) = T+5$. We classify galaxies with $T < -2.5$ as ellipticals,
those with $-2.5 < T < 0.92$ as S0s, and those with $T > 0.92$ as spirals and
irregulars. Galaxies without any bulge are classified as type $T = 9$.
\subsection[]{Selection of IfEs}\label{Selection}
There are a few different selection criteria that are being used to define
isolated field elliptical galaxies in observations \citep[see
e.g.][]{Colbert:2001p574, Marcum:2004p572, Reda:2004p502, Collobert:2006p580}.
Despite the differences in selection criteria, all studies identifying IfEs share
a common ideology. In general, IfEs are defined as elliptical galaxies that do
not have nearby optically bright companion galaxies, usually in the $B-$band.
This is tested by using an isolation sphere or a cone and choosing a minimum
magnitude difference between the brightest and the second brightest galaxy
$\Delta m_{12}$ inside the sphere or the cone. In observational studies, the
isolation cone is often taken as a circle in the sky that expands in redshift
space. Due to peculiar velocities, accurate line-of-sight distances of galaxies
are unknown complicating the identification. Thus, observational studies differ
in the numerical values of the isolation criteria. Despite these differences,
discussed next, we use all possible available observational data for comparison
as the number of observed targets is small. For completeness, we show in Section
\ref{S:Abundance} how the number of IfEs depends on the selection criteria.
\citet{Colbert:2001p574} adopted rather strict rules for isolation, as they
required that IfEs cannot have catalogued galaxies with known redshifts within a
projected radius of $1.0 h^{-1}$Mpc and a velocity of $\pm 1000$ km s$^{-1}$,
while for the morphology of the IfEs they required Hubble types $T \leq -3$.
However, as their source of galaxies was the Third Reference Catalogue of Bright
Galaxies \citep[RC3,][]{1991trcb.book.....D}, their sample IfEs might have faint
companion galaxies due to the incompleteness of the source catalogue. Missing
companions do not affect the study of bright isolated galaxies per se, but
their environment is affected by the incompleteness and missing companions.
\citet{Marcum:2004p572} based their initial sample of IfEs on the same catalogue.
They adopted an even stricter criterion in projected radius, a minimum projected
physical distance of $2.5$ Mpc to any nearest neighbour brighter than $M_{V} = -
16.5$. \citet{Reda:2004p502} used less strict criteria in their study and
required only that an IfE candidate has no neighbours with brightness difference
$\Delta m_{12} \leq 2.0$ mag, in the $B-$band, within $0.67$ Mpc in the plane of
sky and $700$ km s$^{-1}$ in recession velocity. Further, their elliptical
galaxies are also of Hubble type $T \leq -3$.
For the selection of IfEs from the Millennium Simulation we adopt the criteria
used by \citet{Smith:2004p250}, similar to those adopted by
\citet{Reda:2004p502}
We concentrate on galaxies that are incontrovertibly ellipticals
and thus adopt a strict morphology limit $T \leq -4$. We further limit the brightness
of the IfE candidates with the magnitude cut-off of $M_B\leq -19$.
For the isolation criterion we adopt two isolation spheres with the radii of $0.5h^{-1}$
and $1.0 h^{-1}$Mpc. Within these isolation spheres, we require that the $B-$band
magnitude difference of the first- and second-ranked galaxies $\Delta m_{12}$
must be $\geq 2.2$ and $\geq 0.7$ mag, respectively. The application of the
criteria is illustrated in \citet[][Fig.$1$]{Smith:2004p250}.
Our adopted values of magnitude differences correspond to factors of about eight
and two in luminosity for the small and large sphere, respectively. This choice
ensures that possible companions are small and light enough not to produce any
major perturbations in the gravitational potential of the system. Note, that in
case of simulations our isolation criteria operate in real space rather than in
redshift space. Therefore, we can use isolation spheres rather than cones and we
are not plagued by interlopers due to inaccuracies in distance measurements.
\subsection[]{Simulated IfEs and Control Samples}
To identify isolated field elliptical candidates in the MS, we chose five
independent cubic volumes inside the simulation box. Initially, we chose each
volume to have a side length of $\sim 200h^{-1}$ Mpc, while none of the cubes
overlap. However, we later limited the volume from where the possible IfE
candidates can be identified to have a side length of $195h^{-1}$ Mpc. This was
done to enable the study of the surroundings of each IfE candidate in the same
fashion. Moreover, without the limitation, we could have accidentally identified
a field elliptical candidate $\leq 1.0 h^{-1}$Mpc from an edge of the box,
leading to a possible false identification. The five volumes were chosen to
overcome computational issues and for easier study of IfE environment. With five
independent volumes it is also possible to make comparisons between the IfEs of
each volume.
At first all galaxies inside each volume are treated as possible IfE candidates.
After identifying all candidates with $T \leq -4$ and $M_{B} \leq -19$ we apply
the isolation criteria discussed in the previous Section. This produces an
initial list of $302$ galaxies that fulfill the criteria for isolated field
elliptical galaxies. After compiling the initial list of IfEs, we inspected every
galaxy individually, and found that nine candidates were not the main galaxy of
the dark matter halo they reside in. Therefore, all galaxies that fulfill the
observationally motivated IfE criteria are not the central galaxies of their
Friends-of-Friends dark matter groups. In a strict sense these are not isolated
galaxies, as they belong to a dark matter halo containing galaxies more massive
and more luminous than the IfE candidate. Therefore, we omit them from the final
list of IfEs, which contains $293$ galaxies.
We also compile two control samples of elliptical galaxies for comparisons. The
first control sample contains in total $4563$ elliptical ($T \leq -4$) galaxies
brighter than $-19$ in the $B-$band at the redshift $z = 0$ and it is named as
Ellipticals (abridged as Es). For the second control sample, called Main
Ellipticals (abridged as MEs), we only select galaxies of the first control
sample that are the main galaxies of their dark matter haloes. This requirement
further limits the Ellipticals sample, and we are left with $1209$ galaxies in
the Main Ellipticals sample.
\subsubsection{Non-standard IfEs}
The removed galaxies are interesting from another point of view than regular
IfEs, as these nine non-standard IfEs are 'subhalo' galaxies (i.e. satellites of
a larger galaxy) and have multiple nearby companions, from $30$ to $80$ inside
the large isolation sphere. Thus, these galaxies are not isolated in galaxy
number density, but reside in a cluster rather than in the field. Physically it
is obvious that a satellite galaxy is not the dominant galaxy in its dark matter
halo. Since we find these non-standard IfEs, our observationally motivated
isolation criteria are not strict enough and the radii of the isolation spheres
should be slightly larger to avoid any false identifications. But as the number
of non-standard IfEs is small, we can keep the criteria and remove the
non-standard galaxies from our final sample, as described above.
The satellite non-standard ``isolated'' elliptical galaxies can shed some light
on the observational result \citep{Smith:2008p627} where an IfE galaxy (NGC 1600)
was also found to be located off from the dynamical centre of the system.
\cite{Sivakoff:2004p628} found that the X-ray emission is centered slightly to
the northeast of NGC 1600, suggesting that the galaxy is not at the centre of the
gravitational potential. The dynamical study and the X-ray observations together
suggest that NGC 1600 is surrounded by a massive halo extending out to several
hundred kiloparsecs. A subhalo galaxy, fulfilling the optical IfE criteria, could
explain the observations of NGC 1600. All nine galaxies in our simulations reside
inside a large dark matter halo, and in observations would be found to be off
from the centre of the potential well, as in the case of NGC 1600. However, a
massive and more luminous galaxy than NGC 1600 has not been detected at the
centre of the potential well in observations, while this is often the case in
simulations. Thus, a more straightforward explanation, where the IfE orbits
around the centre of the potential well and therefore seems to be shifted from
the centre, seen in X-ray observations, is more plausible. This is in agreement
with an X-ray study of another isolated elliptical NGC 4555, which is assumed not
to lie in the centre of a massive group-scale dark matter halo
\citep{OSullivan:2004p569}.
\section[]{Properties of observed and simulated field elliptical
galaxies}\label{Results}
Below we compare the properties of simulated IfEs with those of the observed
IfEs. If these are close enough, it will justify the theoretical study of the
formation, evolution and merger histories of IfEs, using the simulated IfE
population. Due to the large number of simulated IfEs we can actually make
theoretical predictions for properties, formation mechanisms and evolution of
IfEs that can be tested against observational data when the sample of
observations is large enough.
In following sections we use the Kolmogorov$-$Smirnov (KS) two-sample test to
study the maximum deviation between the cumulative distributions of two samples.
The null hypothesis of the KS test is that the two samples are from the same
population. We give our results as probabilities (p-values) that the difference
between the two samples could have arisen by change if they are drawn from the
same parent population.
\subsection[]{Number of IfEs}\label{S:Abundance}
The fraction of isolated field elliptical ($T \leq -4$ and $M_{B} \leq -19$ mag)
galaxies among all galaxies of any brightness, morphology or dark matter halo
status in the MS is merely $\sim 3.5 \times 10^{-3}$ per cent, corresponding to a
number density of $\sim 8.0 \times 10^{-6}$h$^{3}$ Mpc$^{-3}$. This is a very low
number density, however, such a straightforward calculation is not totally
justified. More meaningful number to conside is the fraction of IfEs among
elliptical galaxies. Moreover, our morphology limit ($T \leq -4$) for a simulated
IfE is very strict, biasing the fraction of elliptical galaxies to significantly
lower fractions than observed in the real Universe. Thus, we also identified
isolated field elliptical galaxies with relaxed morphology limit $T < -2.5$ and
quote the fractions below.
About $0.19$ per cent of all simulated elliptical (now $T < -2.5$) galaxies of
any brightness, can be identified as IfEs, when the criteria of Section
\ref{Selection} are adopted for IfEs. If we relax the strict morphology limit of
the IfEs and require that IfEs also have $T < -2.5$ the fraction rises to $\sim
0.48$ per cent as about $2.5$ times more IfEs can be identified. Relaxing the
morphology limit will therefore change the number density of IfEs to $\sim 1.9
\times 10^{-5}$h$^{3}$ Mpc$^{-3}$. When the brightness of the simulated
elliptical galaxies and their morphologies are limited to $M_{B} \leq -19$ at $z
= 0$ and $T \leq -4$, respectively, about $6.4$ per cent of these galaxies are
identified as IfEs at redshift zero. If we further limit our simulated elliptical
galaxies to galaxies that are the main galaxies of their dark matter haloes (as
in case of IfEs), $\sim 32$ per cent of ellipticals are now identified as IfEs.
Thus, IfEs are very rare objects among all galaxies, but at the same time, over
$30$ per cent of main elliptical galaxies brighter than $-19$ mag in $B-$band can
be classified as IfEs.
\cite{Stocke:2004p575} found from magnitude limited ($m_{B} \leq 15.7$)
observations that less than three per cent of all galaxies can be classified as
isolated. This value should be interpreted as an upper limit for the number of
IfEs, as it was based on all types of galaxies. \cite{Stocke:2004p575} further
found that early-type galaxies outnumber S$0$ galaxies $2:1$ in very isolated
areas. However, as spiral galaxies are the most dominant galaxy type in low
densities \citep[e.g.][]{Dressler:1980p645}, the observable estimate of the
number of IfEs is significantly lower than one per cent.
\citet{HernandezToledo:2008p648} concluded that early-type galaxies amount to
only $3.5$ per cent among all isolated galaxies, lowering the observational
estimate of IfEs to $\sim 1 \times 10^{-2}$ per cent. This fraction is in good
agreement with our findings as about $\sim 9 \times 10^{-3}$ per cent of all
simulated galaxies can be identified as IfEs when $T < -2.5$ morphology limit has
been adopted.
As observational studies adopt different criteria for isolation, we illustrate
their influence on the number of IfEs, with a stricter set of parameter values.
For simplicity, we adopt only one isolation sphere with a radius of $2.5h^{-1}$
Mpc, and require that the magnitude difference between the brightest and the
second brightest galaxy $\Delta m_{12}$ is $\geq 2.0$ mag inside the isolation
sphere. For the Hubble type, we adopt the same requirement as previously for our
IfEs: $T \leq -4$. Such values were used, e.g., by \citet{Marcum:2004p572}. With
these values we find that only $\sim 3 \times 10^{-4}$ per cent of all MS
galaxies can be classified as IfEs, corresponding to an extremely low number
density of $\sim 6 \times 10^{-7}$h$^{3}$ Mpc$^{-3}$. This is approximately ten
times less than before. It is obvious that changing the isolation criteria has a
major impact on the number of IfEs and therefore their statistical properties, as
stricter parameter values require a galaxy that is located in a true void. This
result also shows that galaxies with comparable luminosities tend to reside in
relatively close proximity rather than being truly isolated from any other
reasonable sized companion.
The results of this section show that the fraction of isolated field elliptical
galaxies depends highly on the isolation criteria adopted, but also on the
morphological type that is required. As the morphological type of simulated
galaxies is less than precise our number density for IfEs should not be taken as
a strict limit, but as a guideline when observations are being planned. Thus,
higher fractions of IfEs can be expected if isolation criteria or morphology
limits are being relaxed.
\subsection[]{Colour-magnitude Diagrams}\label{Comparison}
The easiest property of IfEs to observe must be the luminosity, and thus the
colour-magnitude relation. This simple yet powerful relation is studied in Figs.
\ref{colourMag} and \ref{colourMag2}, which show the colour-magnitude diagrams
(CMDs) of simulated and observed IfEs. Fig. \ref{colourMag} shows a comparison
between the simulated IfEs and observed field elliptical galaxies presented in
\citet{Reda:2004p502}, while Fig. \ref{colourMag2} shows a comparison to the
sample of IfEs presented in \citet{Marcum:2004p572}. We also plot the Ellipticals
(our control sample) for completeness. Note, however, that in this section we
limit the discussion of the colour-magnitude diagrams to the comparison of
simulated and observed IfEs, while the differences between simulated IfEs and Es
are discussed in Section \ref{ComparisonA}. Colours and associated errors and
standard deviations of observed and simulated IfEs and simulated elliptical
galaxies (Es) are listed in Table \ref{tb:colours}.
Fig. \ref{colourMag} shows that the observed IfEs lie in slightly bluer region in
the CMD than simulated IfEs of the same brightness and that they seem to follow a
tighter correlation than simulated IfEs. The $B - R$ colour scatter of simulated
IfEs is slightly larger compared to observed IfEs: all simulated IfEs show a
$1\sigma$ deviation of $0.23$, however, as the observations of
\citet{Reda:2004p502} are limited to relatively bright ($M_{R} \leq -21.5$)
galaxies, we calculate the colour scatter of bright simulated IfEs using
magnitude cut-off of $M_{R} \leq -21.5$ and find it to be $0.17$. The colour
scatter of the bright simulated IfEs is close to the observed $B - R$ colour
scatter ($0.16$), yet slightly higher. However, one should bear in mind that the
number of observed IfEs is still small.
\begin{table}
\caption{Colours of simulated and observed IfEs and simulated non-isolated
elliptical galaxies (Es).}
\label{tb:colours}
\begin{tabular}{lcclc}
\hline
Sample & $\Delta mag$ & Mean Colour & Error & $\sigma$\\
\hline
IfEs & $B - R$ & $1.47$ & $0.01$ & $0.23$ \\
IfEs$^{1}$ & $B - R$ & $1.57$ & $0.01$ & $0.17$ \\
IfEs & $B - V$ & $0.79$ & $0.01$ & $0.15$ \\
IfEs$^{2}$ & $B - V$ & $0.82$ & $0.01$ & $0.14$ \\
Es & $B - R$ & $1.58$ & $0.002$ & $0.10$ \\
Es & $B - V$ & $0.86$ & $0.01$ & $0.07$\\
\citet{Reda:2004p502} & $B - R$ & $1.49$ & $0.06$ & $0.16$\\
\citet{Marcum:2004p572} & $B - V$ & $0.76$ & $0.06$ & $0.18$\\
\hline
\end{tabular}\\
\medskip{Note: The error refers to the standard error of the mean while $\sigma$
is the standard deviation. IfEs$^{1}$ and IfEs$^{2}$ refers to samples where
faint simulated isolated field elliptical galaxies have been removed and the
IfE samples contain only galaxies with $M_{R} \leq -21.5$ and $M_{B} \leq
-19.5$, respectively.}
\end{table}
\citet{Reda:2004p502} found a mean effective colour of $(B - R)_{e} = 1.49 \pm
0.06$ for their isolated elliptical galaxies, which is in good agreement with our
mean value of $1.47 \pm 0.01$ (the errors are the standard error on the mean, see
Table \ref{tb:colours} for details). If we limit our sample of simulated IfEs to
galaxies brighter than $M_{R} \leq -21.5$, as in the sample of
\citet{Reda:2004p502}, the mean $B - R$ colour becomes $1.57 \pm 0.01$. This is
redder than the mean colour in \citet{Reda:2004p502} and shows that our mean
colour is significantly affected by faint IfEs of the blue cloud. When the KS
test is applied to the observed \citep{Reda:2004p502} and all simulated IfE
colours the p-value ($\sim 0.33$) shows that we cannot reject the null hypothesis
at $30$ per cent level. However, if we consider only bright ($M_{R} \leq -21.5$)
simulated IfEs, the p-value is only $\sim 0.01$, implying that a difference
between the observed and simulated IfE $B - R$ colours exists when faint
simulated IfEs are excluded from the comparison.
A straigth line fit to both observed and simulated IfEs differs in the slopes and
intercepts (see Fig. \ref{colourMag}). The linear fit of observed IfEs shows a
steeper slope than the fit of the simulated IfEs. However, as the IfE sample of
\citet{Reda:2004p502} miss galaxies fainter than $M_{R} > -21.5$ it is not
entirely clear what the order of magnitude of this difference might be, and due
to the small number of observed IfEs in the sample of \citet{Reda:2004p502} only
a single new data point at the faint end of the CMD could change the fit
significantly.
The $V-$band (Fig. \ref{colourMag2}) CMD differs from the $R-$band diagram as
correlations seem significantly looser. Here the colour scatter of the observed
IfEs is larger than for the simulated IfEs ($1\sigma$ deviations are $0.18$ and
$0.15$, respectively). The IfE sample of \citet{Marcum:2004p572} extends to
slightly fainter magnitudes than the sample of \citet{Reda:2004p502}, but their
fainter IfEs are surprisingly red ($B - V \sim 0.9$). The simulated IfEs show
slightly redder colours than observed IfEs; the mean $B - V$ colours are $0.79
\pm 0.01$ and $0.76 \pm 0.06$, for simulated and observed IfEs, respectively. The
colours of simulated IfEs agree with observed colours within their standard
errors of the mean, and the KS test approves the null hypothesis with high
probability (p-value $\sim 0.63$) when colours of the simulated IfEs are compare
to the colours of the IfE sample of \citet{Marcum:2004p572}. The p-value would be
only $0.04$ if the observed IfEs of \citet{Marcum:2004p572} were compared against
the simulated Es. Thus, the simulated IfEs agree well with the observed IfEs of
\citet{Marcum:2004p572} when $B - V$ colour samples are compared.
Figs. \ref{colourMag} and \ref{colourMag2} show a separate population of
simulated IfEs that belong to the blue cloud and populate the faint end of the
CMD. Quantitatively the colour and the brightness of this population is
following: $B - R \leq 1.4$ or $B - V \leq 0.7$ and $M_{R} > -21.5$ or $M_{B} >
-20.0$. The population of faint and blue IfEs comprise $\sim 26$ per cent ($76$)
of all simulated IfEs, thus, every fourth IfE belong to this population. We note
that none of the IfEs of \citet{Reda:2004p502} or \citet{Marcum:2004p572}
populates this faint and blue part of the CMDs. The sample of
\citet{Marcum:2004p572} shows two very blue IfEs that are part of the global blue
cloud, but are not part of the population of faint and blue IfEs that is found
from the simulation. However, this might be due to the magnitude limit of
\citet{Marcum:2004p572} sample, as they do not have IfEs fainter than $M_{B} >
-19.5$.
\citet{Marcum:2004p572} found preliminary evidence that $50$ per cent of their
sample of isolated early-type galaxies show blue global colours. However, because
of the small sample size they could not conclude that the higher occurrence of
blue E-type galaxies is related to the extremely low densities of the associated
environments. As we are not limited by a small sample, we can confirm whether
IfEs show bluer global colours than non-isolated early-type galaxies or not. The
mean $B - V$ colour of simulated IfEs is $\sim 0.79 \pm 0.01$ in agreement with
the findings of \citet{Marcum:2004p572}. However, this value is significantly
affected by the population of faint IfEs; if we remove the faint IfEs and
consider only IfEs with $M_{B} \leq -19.5$ the mean colour changes to $\sim 0.82
\pm 0.01$ (see Table \ref{tb:colours} and also Fig.~\ref{f:colour} of the next
Section). This is close, yet slightly bluer, than the mean $B - V$ colour of
simulated ellipticals $(0.86 \pm 0.01)$. A similar result is seen when the $B -
R$ colours are studied; bright $(M_{R} \leq -21.5)$ simulated IfEs are almost as
red as all simulated non-isolated ellipticals (Es) $ B - R \sim 1.57 \pm 0.01$
and $1.58 \pm 0.002$, respectively. Thus, our results show that simulated IfEs do
not show global blue colours if faint IfEs are removed, but the blue colours are
due to the separate, faint and blue population of IfEs. Our theoretical findings
predict that $\sim 26$ per cent of IfEs show global blue colours and that these
IfEs belong to the separate population of faint and blue IfEs.
\begin{figure}
\includegraphics[width=84mm]{ColMagRelation.eps}
\caption{Colour ($B-R$) versus absolute $R-$band magnitude for the observed and
simulated IfEs. The data for the Es control sample have been plotted for
comparison. For IfEs and Es a linear fit is shown. For clarity, only every 10th
Es has been plotted. Note the separate population of faint and blue IfEs.}
\label{colourMag}
\end{figure}
\begin{figure}
\includegraphics[width=84mm]{ColMagRelation2.eps}
\caption{Colour ($B-V$) versus absolute $B-$band magnitude for the observed and
simulated IfEs. The data for the Es control sample have been plotted for
comparison. For Es a linear fit is shown. For clarity, only every 10th Es has
been plotted. Note the separate population of faint and blue IfEs.}
\label{colourMag2}
\end{figure}
Different trends in the CMDs, visible in Figs. \ref{colourMag} and
\ref{colourMag2}, may be interpreted as environment effects of galaxy formation.
However, it is also possible that the bluest IfEs have had different formation
mechanism and evolutionary path than redder IfEs. Different evolutionary paths
and merging histories could explain different properties observed at redshift
$z=0$, as well as the influence of environment. Both possibilities will be
explored in detail later.
\subsection[]{Dark Matter Halo Masses}\label{R:masses}
We continue comparisons of simulated and observed IfEs by studying their dark
matter haloes. Even it is far from simple task to derive reliable dark matter
halo masses from observations the comparison is highly interesting, as the dark
matter halo properties and galaxy properties are tightly linked.
Isolated field elliptical galaxies in the MS are mainly found residing inside
dark matter haloes that are lighter than $7 \times 10^{12}h^{-1}$M$_{\odot}$. The
dark matter haloes of simulated IfEs are surprisingly light; the median mass is
only $\sim 1.2 \times 10^{12}h^{-1}$M$_{\odot}$. Even the most massive dark
matter halo hosting an IfE is lighter than $2.2 \times 10^{13}h^{-1}$M$_{\odot}$,
which is comparable to a dark matter halo of a small group.
\cite{Memola:2009p642} calculated the total masses of two of their isolated
ellipticals NGC 7052 and NGC 7785 from X-ray observations and quote values $\sim
5 \times 10^{11}$M$_{\odot}$ and $\sim 1.9 \times 10^{12}$M$_{\odot}$,
respectively. These mass values agree well with our findings of dark matter halo
mass. \cite{Norberg:2008p652} did not find any isolated systems residing in
haloes outside the range $\sim 5 \times 10^{11}$ to $10^{13}h^{-1}$M$_{\odot}$ in
their simulations, therefore, we can conclude that IfEs reside in lighter than
$\sim 2 \times 10^{13}h^{-1}$M$_{\odot}$ dark matter haloes.
The dark matter halo mass also sets constrains to the possible formation
mechanisms; IfEs residing in light haloes cannot be merger remnants of groups or
clusters. Only those IfEs that have a large and massive dark matter halo could
have formed via merger of a group or multiple larger galaxies, but even then the
group should have been poor with only few member galaxies. Evidently, this is
true for fossil groups which are in many ways similar objects to IfEs. The
dynamical masses of fossil groups range from $\sim 10^{13}$ to
$10^{14}h^{-1}$M$_{\odot}$ \citep[see e.g.][]{DiazGimenez:2008p617}.
\subsection[]{Ages}\label{R:ages}
The last property of IfEs we compare is their age. The median mass weighted age
of our model IfEs is $8.84$ Gyr (see Table \ref{tb:properties}). We also note
that the number density of young (mass weighted age less than $5$ Gyr) IfEs is
extremely low (Fig. \ref{f:masswAge}), giving a lower limit for the age of an
IfE.
\citet{Reda:2005p561} found a mean age of their IfEs to be $4.6 \pm 1.4$ Gyr,
while \citet{Proctor:2005p558} quote an age estimate of $\sim 4$ Gyr. These age
estimates are around the young end of our values. However, these estimates should
not be compared directly to our values as the definitions are different: mass-
vs. luminosity-weighted age. Moreover, \cite{Collobert:2006p580} found a broad
range of stellar ages for their IfEs; ranging from $\sim 2$ to $15$ Gyr, in good
agreement with our estimates, as the mass weighted age of model IfEs ranges from
$5$ to $12$ Gyrs. The big scatter suggests that the formation and evolution of
IfEs is not concentrated at a fixed epoch. This can also explain the large
scatter we see in some properties of IfEs, as different formation times and
evolutionary paths can lead to significantly different properties at $z = 0$.
\subsection[]{Environments}\label{R:environment}
To get a complete picture of the properties of isolated field elliptical
galaxies, we have to look at their environment. We define a companion as a galaxy
that resides inside either of the two isolation spheres; the small $0.5h^{-1}$
Mpc or the large sphere $1.0h^{-1}$ Mpc. If not specifically mentioned, the
number of companions refers to the total number of galaxies, $N_{COMP}$, inside
the large isolation sphere. Our definition of a companion does not guarantee that
the galaxy belongs to the same dark matter halo as the IfE. Thus, it is possible
that some of the companion galaxies are not gravitationally bound to the IfE.
Truly isolated galaxies, without any companions in simulations, should be treated
with caution. It may be that the mass resolution is not sufficient to form small
subhaloes that could harbour a dwarf galaxy.
IfEs can have from $0$ to $\sim 30$ companion galaxies. We do not find a single
IfE with more than $30$ dwarf companions, indicating that our IfEs are in
relatively low density environments. The majority of IfEs have $0$ to $20$
companions, while the mean value of the number of companions, $N_{COMP}$, is 10.7
in the MS. For IfEs with the total number of companion galaxies less than $17$,
most of the companions are found to reside inside the smaller ($0.5h^{-1}$ Mpc)
isolation sphere. As the magnitude difference limit inside the small sphere is
$2.2$ we can be sure that our IfEs (excluding the nine subhalo galaxies discussed
earlier) are well isolated from bright nearby galaxies. So, IfEs do not reside
inside cluster-sized dark matter haloes, with virial radii $\sim 1.5 h^{-1}$ Mpc,
as most their companion galaxies are within $0.5h^{-1}$ Mpc distance from the
halo's main galaxy. If IfEs resided in cluster haloes, we should find companion
galaxies also in the larger isolation sphere. We would also expect to find a
larger number of dwarf companions.
\citet{Reda:2004p502} found that only the very faint dwarf galaxies ($M_{R}
\geq -15.5$) appear to be associated with isolated ellipticals.
On the contrary, we find companion galaxies with a broad range of magnitudes from
the MS. The mean $R-$band magnitude for these companions is $-17$ mag. The
quartile values for companion $R-$band magnitudes are $-18.2$ and $-15.8$ mag showing that
model IfEs can have relatively bright companion galaxies. Thus, the magnitude gap
between companion galaxies and an IfE seems to be larger in observations than in
simulations. This may be due to the way how galaxy luminosities are treated in
the SAM. Thus, this result is probably not without a bias due to the limiting
mass resolution of the MS.
In Fig. \ref{f:companionDistance} we show the distribution of companion galaxy
distances from IfEs. We have divided the IfEs sample into two, to separate the
population of blue, light and faint IfEs; the IfE sample is divided by dark
matter halo mass. Fig. \ref{f:companionDistance} clearly shows that more massive
IfEs have close companions more often than light IfEs. The mean virial radius of
the dark matter haloes of heavy IfEs ($M_{DM} > 10^{12}$M$_{\odot}$) is
$225h^{-1}$ kpc, while it is $115h^{-1}$ kpc for light IfEs ($M_{DM} \leq
10^{12}$M$_{\odot}$). A significant number of companions of heavy IfEs are inside
the mean virial radius and belong to the same dark matter halo as the IfE.
However, for blue, light and faint IfEs the trend is opposite; we find most of
the companions more than six times the mean virial distance away from the IfE.
The median companion distances are $0.33$ and $0.73h^{-1}$ Mpc for heavy and
light IfEs, respectively.
Similar results are also found if the IfEs are divided by their colour. The blue,
$B-R \leq 1.4$, IfEs have most of their companion galaxies more than five times
away than the mean virial radius, while the red IfEs have the majority of their
companions within the mean virial radius. These results show that the most
isolated field elliptical galaxies are blue and relatively faint.
\begin{figure}
\includegraphics[width=84mm]{massDistance.eps}
\caption{Total number of companion galaxies, $N_{COMP}$, inside the large
isolation sphere for light and massive IfEs.}
\label{f:companionDistance}
\end{figure}
\section[]{Comparison of Regular and Field Ellipticals}\label{ResultsA}
Here we compare the properties of the simulated isolated field elliptical
galaxies to the control sample of all elliptical galaxies in the simulation, to
find out the differences between the two populations. For many properties the
median values are relatively close to each other, but clear differences exist in
the shapes of the distributions. Thus, throughout the following sections we use
the Kolmogorov$-$Smirnov two-sample test to assess whether two distributions are
drawn from the same parent population. Results of the KS tests are given as
probabilities (p-values) and presented in Table \ref{tb:propertiesKSresults}. For
completeness, we also plot the histograms of studied properties (see Figs.
\ref{f:magb} $-$ \ref{f:masswAge}) and present the quartile values in Table
\ref{tb:properties}.
\subsection[]{Morphology}\label{Morph}
The distributions of morphologies ($T$ values) of isolated field elliptical
galaxies and Es are very similar, with median values: $-5.62$ and $-5.80$,
respectively. The KS test approves the null hypothesis with high probability
(p-value $\sim 0.39$), thus, it is likely that the differences in morphology
distributions have arisen by change. This is no surprise as our definition of
IfEs and Es requires that the morphology value is smaller than four.
\subsection[]{Galaxy Colour-magnitude Diagrams}\label{ComparisonA}
The galaxy colour-magnitude diagrams show three main features: the red sequence,
the green valley, and the blue cloud. In general, elliptical galaxies populate
the area known as the red sequence. This is certainly true for most elliptical
galaxies of the MS, however, is this true for IfEs as well?
Figs. \ref{colourMag} and \ref{colourMag2} show that the red sequence of IfEs
starts at $M_{R} \sim -21$ (or $M_{B} \sim -19.5$), while it extends to fainter
magnitudes for Es. A trend of redder colours with increasing luminosity is noted
for both simulated IfEs and Es. However, the slope of the trend is steeper and
the scatter is higher for IfEs. Simulated IfEs show a broader distribution in
colours (Figs. \ref{colourMag} and \ref{colourMag2}) than simulated Es. The
$1\sigma$ scatter of $B - R$ colours is $\sim 0.23$ and $\sim 0.10$ for IfEs and
Es, respectively. For $B - V$ colours the $1\sigma$ values are $\sim 0.15$ and
$\sim 0.07$ for IfEs and Es, respectively. Spearman rank order correlation test
shows a very strong (correlation coefficient $cc \sim -0.65$) correlation for
IfEs (in Fig. \ref{colourMag}) and significant ($cc \sim -0.33$) correlation for
Es. The trends in Fig. \ref{colourMag2} are not as strong according to the
Spearman test: $cc \sim -0.60$ and $\sim -0.27$ for IfEs and Es, respectively.
However, all correlations are highly significant, as the probability to have as
large correlation coefficients for uncorrelated data is $< 10^{-20}$ in all
cases.
The colour correlations in Figs. \ref{colourMag} and \ref{colourMag2} suggest
that IfEs follow a different colour-magnitude trend than Es. The KS tests show
large differences (D-values) when the colours of IfEs are compared to Es. The KS
test rejects the null hypothesis in case of IfEs and Es $B - R$ and $B - V$
colours with high probability (p-values $< 10^{-15}$ in both cases). If we
exclude the separate population of faint and blue IfEs, the previous statement
does not change, only the probabilities (p-value now $< 10^{-6}$). Thus, the
CMDs of IfEs and Es differ significantly even when the separate population of
faint and blue IfEs is removed.
\subsection[]{Colour-mass Diagrams}\label{R:cmassA}
Fig. \ref{f:colourDmMass} shows the $B - R$ colour of a galaxy as a function of
the mass of the underlying dark-matter halo. We note from Fig.
\ref{f:colourDmMass} that the Es show a rather constant relation with a few
outliers; in general, Es have the $B-R$ colour $\sim 1.6 \pm 0.15$ independent of
the mass of the dark matter halo they reside in. However, a completely different
trend is seen for IfEs, as they tend to get redder when the dark matter halo mass
grows, while light ($M_{DM} < 10^{12}h^{-1}$ M$_{\odot}$) dark matter haloes host
blue IfEs. The $B-R$ colours of IfEs grow steeply as a function of dark matter
halo mass below $10^{12}h^{-1}$M$_{\odot}$ indicating the influence of dark
matter halo properties to the IfE galaxy they host.
All IfEs are the main galaxies of their dark matter haloes. If instead of the Es
sample we used the MEs for comparison, Fig. \ref{f:colourDmMass} would change.
There would still be a significant number of red, $B-R \sim 1.6$, non-isolated
elliptical galaxies residing in dark matter haloes lighter than $M_{DM} <
10^{12}h^{-1}$ M$_{\odot}$, but their total number would be significantly
smaller. MEs that reside in light ($M_{DM} < 10^{12}h^{-1}$ M$_{\odot}$) dark
matter haloes show a big scatter in colours, indicating that not only the
dark matter halo, but also the environment and the formation history of the halo
effects the galaxy it hosts.
\begin{figure}
\includegraphics[width=84mm]{ColMass.eps}
\caption{Dark matter halo mass versus colour ($B-R$) of the galaxy. For clarity,
only every 5th Es has been plotted.}
\label{f:colourDmMass}
\end{figure}
Fig. \ref{f:colourDmMass} further shows that the population of faint and blue
IfEs noted in the colour-magnitude diagrams resides in the lightest dark matter
haloes with $M_{DM} < 10^{12}h^{-1}$ M$_{\odot}$. Below this mass scale we see a
strong correlation between the dark matter halo mass and the evolution of the
galaxy. Thus, simulations predict an unobserved population of IfEs that are blue
and faint and that reside in light dark matter haloes.
\subsection[]{Colour and Luminosity Distributions}\label{R:coloursA}
Fig. \ref{f:magb} shows the number density of absolute $B-$band rest frame
magnitudes for IfEs and Es. The distributions look very different and the KS test
(Table \ref{tb:propertiesKSresults}) rejects the null hypothesis with high
probability. The differences in the luminosity functions are likely due to the
isolation criteria of IfEs. The most interesting result, seen in the Figure, is
that isolated field elliptical galaxies have an almost constant, $\sim 8 \times
10^{-6}$, number density throughout their $B-$band magnitudes. Surprisingly, when
only brighter ($ -21.7< $ M$_{B} < - 21.0$) elliptical galaxies are considered,
it is almost as probable to find an isolated field elliptical as it is to find a
non-isolated elliptical. It is also noteworthy that we have not identified a
single IfE galaxy brighter than $-21.7$ mag (in $B-$band), while the brightest Es
are almost one magnitude brighter ($\sim -22.5$). This result is natural, since
the brightest elliptical galaxies are normally found to reside in centres of
large clusters. The brightest cluster galaxies (BGCs) have M$_{V} \sim -23.5$ mag
and they have usually experienced many merging events during their evolution
\citep[e.g.][]{DeLucia:2007p414}. Multiple merging events have brought more mass
and gas to the central galaxy and have caused massive star formation and greater
luminosity. It is unlikely that IfEs have followed the same evolutionary path
(see Fig. \ref{f:dmMass} and \ref{f:stMass}).
\begin{figure}
\includegraphics[width=84mm]{magb.eps}
\caption{Number density of B$-$band rest frame magnitude for isolated field
ellipticals (IfEs) and Es.}
\label{f:magb}
\end{figure}
Since the $B-$band luminosity function of IfEs is almost constant, it leads to
another question, namely if IfEs have a constant number density in colour as
well. Fig. \ref{f:colour} replies to this question and presents the number
density for the B$-$R colour. The distributions for Es and IfEs look very
different and the KS test (Table \ref{tb:propertiesKSresults}) quantifies that
these distributions are not drawn from the same parent population. The isolated
field elliptical galaxies show a bimodal distribution while Es show clearly only
one peak. This result is also seen when the values of first quartiles of both
distributions (Table \ref{tb:properties}) are compared. The value of the first
quartile of our IfEs is almost $0.3$ magnitudes bluer than for the Es in
agreement with \citet{Marcum:2004p572} who found evidence that $50$ per cent of
their sample of isolated early-type galaxies show blue global colours.
\begin{figure}
\includegraphics[width=84mm]{colour.eps}
\caption{Number density of colour distribution ($B-R$ magnitudes) for isolated
field ellipticals (IfEs) and Es.}
\label{f:colour}
\end{figure}
The number density of isolated and non-isolated elliptical galaxies is quite
similar for blue, $B-R < 1.3$, galaxies. The bimodality of the IfE
distribution and the blue peak present in Fig. \ref{f:colour} is caused by the
population of blue, faint and light IfEs. The colour distribution of IfEs
suggests that some isolated field ellipticals have either formed later or have
had merger activity at lower redshifts than non-isolated ellipticals. We confirm
whether this is the case in Section \ref{Formation} where formation and merging
times of IfEs are compared to non-isolated ellipticals.
\subsection[]{Dark Matter and Stellar Masses}\label{R:massesA}
We continue our comparisons between IfEs and Es by studying their masses and
composition. Both samples show very little cold gas: $< 1.0 \times 10^{10}h^{-1}$
M$_{\odot}$, with no ongoing star formation (the median star formation rate is
$\sim 0$ M$_{\odot}$yr$^{-1}$). This is typical for elliptical galaxies that are
usually considered as ``red and dead'' at $z = 0$.
When dark matter halo masses are studied in detail, an interesting result is
found. Isolated field elliptical galaxies reside mainly inside dark matter haloes
that are lighter than $7 \times 10^{12}h^{-1}$M$_{\odot}$, while Es are found to
reside more often in more massive haloes. The differences between the dark matter
haloes of Es and IfE galaxies are striking, which is clearly visible in Fig.
\ref{f:dmMass} where we plot the dark matter halo mass distribution as a function
of number density. It is no surprise that the KS test (Table
\ref{tb:propertiesKSresults}) does not approve the null hypothesis, especially as
the distribution of IfEs shows weak bimodality. The great difference in dark
matter halo mass is interesting as the properties of galaxies are tightly related
not only to the environment, but also to the dark matter halo mass within the
galaxy resides in \citep{Croton:2008p600}.
\begin{figure}
\includegraphics[width=84mm]{DMmass.eps}
\caption{Number density of dark matter halo mass for isolated field ellipticals
(IfEs) and Es.}
\label{f:dmMass}
\end{figure}
\begin{figure}
\includegraphics[width=84mm]{stellarMass.eps}
\caption{Number density of stellar mass for isolated field ellipticals (IfEs) and
Es.}
\label{f:stMass}
\end{figure}
While the dark matter haloes of IfEs are, in general, significantly lighter than
haloes of Es, we see a different trend when stellar masses are compared. Even
though the distributions in Fig. \ref{f:stMass} may look somewhat similar, the KS
test (Table \ref{tb:propertiesKSresults}) disproves the null hypothesis with high
probability. The differences in stellar mass distributions become very clear if
we consider the quartile values of both samples. Isolated field elliptical
galaxies have, in general, more stellar mass than Es (Table \ref{tb:properties}).
This is especially intriguing as IfEs reside inside rather light dark matter
haloes (see Fig. \ref{f:dmMass}). To emphasize this difference we plot the dark
matter halo mass versus the stellar mass for both samples in Fig.
\ref{f:massComparison}. IfEs are found to contain more stellar matter in respect
to dark matter than Es. The stellar mass of IfEs grows almost linearly with the
dark matter mass with a mass (dark/stellar) ratio of $\sim 4 \times 10^{-2}$. It
is possible that the difference is related to the formation and evolutionary
paths of IfEs. This possibility is studied in detail in Section \ref{Formation}.
\begin{figure}
\includegraphics[width=84mm]{stellarvsmass.eps}
\caption{A scatter plot of stellar and dark matter halo mass. Yellow diamonds
correspond to the sample of isolated field elliptical galaxies (IfEs), while red
circles correspond to Es. For clarity, only every 2nd galaxy has been plotted.}
\label{f:massComparison}
\end{figure}
\begin{table}
\caption{Mean values of the properties and results of the Kolmogorov$-$Smirnov
(KS) test: isolated field elliptical (IfEs) vs. control sample (Es) galaxies.}
\label{tb:propertiesKSresults}
\begin{tabular}{lrrc}
\hline
Quantity & IfEs & Es & KS Probability (p-value)\\
\hline
Mass$_{DM}$ & $186.29$ & $6751.93$ & $ < 10^{-10}$\\
Mass$_{ST}$ & $5.46$ & $3.92$ & $ < 10^{-10}$\\
Mass$_{CG}$ & $0.32$ & $0.10$ & $ < 10^{-10}$\\
Age & $8.58$ & $9.32$ & $ < 10^{-10}$\\
Colour & $1.47$ & $1.58$ & $ < 10^{-10}$\\
M$_{B}$ & $-20.19$ & $-19.76$ & $ < 10^{-10}$\\
$T$ & $-6.60$ & $-7.06$ & $0.39$\\
\hline
\end{tabular}\\
\medskip{Note: Quantity Mass$_{DM}$ refers to virial dark matter mass of the
background halo, Mass$_{ST}$ refers to stellar mass, while Mass$_{CG}$ refers to
the mass in cold gas. All masses are in units of $10^{10}h^{-1}$M$_{\odot}$. Age
is the mass-weighted age of a galaxy in units of $10^{9}$yr. Colour is $B-R$ in
absolute restframe (Vega) magnitudes, and M$_{B}$ is the absolute restframe
(Vega) magnitude in the $B-$band (Buser B3 filter). $T$ is the morphology, and it
is the only quantity for which the null hypothesis of the KS test is approved.}
\end{table}
\subsection[]{Age distributions}\label{R:agesA}
Distributions of mass weighted age (Fig.~\ref{f:masswAge}) look rather similar,
however, the KS test (Table \ref{tb:propertiesKSresults}) does not approve the
null hypothesis. When quartile values (see Table \ref{tb:properties}) are studied
it is obvious that IfEs have lower mass weighted age than Es, suggesting that
IfEs are younger. Even IfEs seem to be slightly younger than Es in statistical
sense, they cover roughly the same age range, only the extremes are missing.
\begin{figure}
\includegraphics[width=84mm]{massWeightedAge.eps}
\caption{Number density of the mass weighted age for isolated field ellipticals
(IfEs) and Es.}
\label{f:masswAge}
\end{figure}
\begin{table}
\caption{Quartile values of the properties of isolated (IfE) and control sample
(Es) galaxies.}
\label{tb:properties}
\begin{tabular}{lrrrr}
\hline
Sample & Quantity & $1^{st}$ Quartile & Median & $3^{rd}$ Quartile\\
\hline
IfEs & Mass$_{DM}$ & $34.86$ & $120.58$ & $235.48$\\
Es & Mass$_{DM}$ & $311.95$ & $1069.28$ & $4998.61$\\
IfEs & Mass$_{ST}$ & $1.69$ & $4.34$ & $7.94$\\
Es & Mass$_{ST}$ & $1.97$ & $2.84$ & $4.73$\\
IfEs & Mass$_{CG}$ & $0.10$ & $0.22$ & $0.44$\\
Es & Mass$_{CG}$ & $0.04$ & $0.05$ & $0.09$\\
IfEs & Age & $7.21$ & $8.84$ & $10.03$\\
Es & Age & $8.44$ & $9.39$ & $10.41$\\
IfEs & Colour & $1.28$ & $1.59$ & $1.64$\\
Es & Colour & $1.56$ & $1.60$ & $1.63$\\
IfEs & M$_{B}$ & $-20.70$ & $-20.18$ & $-19.64$\\
Es & M$_{B}$ & $-20.14$ & $-19.65$ & $-19.28$\\
IfEs & $N_{COMP}$ & $3.00$ & $7.00$ & $15.00$\\
\hline
\end{tabular}
\medskip{Note: Mass$_{DM}$ is the virial dark matter mass of the background halo,
Mass$_{ST}$ -- the stellar mass, Mass$_{CG}$ -- the mass in cold gas. All masses
are in units of $10^{10}h^{-1}$M$_{\odot}$. Age is the mass-weighted age of a
galaxy in units of $10^{9}$yr. Colour is $B-R$ in absolute restframe (Vega)
magnitudes, and M$_{B}$ is the absolute restframe (Vega) magnitude in the
$B-$band (Buser B3 filter). $N_{COMP}$ is the number of companion galaxies within
the large ($1.0h^{-1}$ Mpc) avoidance sphere around the isolated field elliptical
galaxy.}
\end{table}
\section[]{Formation and Evolution of IfEs}\label{Formation}
The analysis of simulated IfEs in the previous sections shows similarities in
colour-magnitude diagrams and comparable ages, colours, and dark matter halo
masses with observational data. In general, we do not find big discrepancies
between simulations and observations for the properties of IfEs we can compare.
The KS tests between the IfEs and Es do not approve the null hypothesis, except
for morphology. Therefore, IfEs selected with the criteria of Section
\ref{Selection} form a distinct class of objects and are significantly different
from regular elliptical galaxies. As a next step in our analysis we use
simulation data to study evolutionary paths and formation mechanisms of IfEs.
Note that in this section, unless otherwise stated, we use the MEs sample as the
control sample, which contains only those non-isolated elliptical galaxies that
are the main galaxies of their dark matter haloes.
\begin{table*}
\begin{minipage}{150mm}
\caption{Statistics of formation and evolutionary times.}
\label{tb:times}
\begin{tabular}{lcccccccccc}
\hline
Sample & Time & Mean & Median & $1^{st}$ Quartile & $3^{rd}$ Quartile & Mode &
Min & Max & Stdev\\
\hline
IfEs & $z_{i}$ & $0.644$ & $0.408$ & $0.183$ & $0.687$ & $0.242$ & $0.064$ &
$6.200$ & $0.836$\\
MEs & $z_{i}$ & $1.062$ & $0.624$ & $0.242$ & $1.386$ & $0.116$ & $0.020$ &
$5.724$ & $1.089$\\
IfEs & $z_{a}$ & $1.079$ & $0.989$ & $0.564$ & $1.504$ & $1.173$ &
$0.020$ & $5.289$ & $0.710$\\
MEs & $z_{a}$ & $1.303$ & $1.276$ & $0.755$ & $1.766$ & $1.276$ & $0.041$ &
$4.520$ & $0.737$\\
IfEs & $z_{f}$ & $1.520$ & $1.386$ & $0.828$ & $1.912$ & $1.386$ & $0.012$ &
$5.289$ & $0.831$\\
MEs & $z_{f}$ & $1.055$ & $0.687$ & $0.457$ & $1.386$ & $0.564$
& $0.041$ & $6.197$ & $0.893$ \\
MEs$^{*}$ & $z_{f}$ & $3.214$ & $3.308$ & $2.831$ & $3.576$ &
$3.576$ & $1.276$ & $6.197$ & $0.790$\\
IfEs & $z_{l}$ & $0.310$ & $0.208$ & $0.116$ & $0.408$ & $0.116$ &
$0.020$ & $2.070$ & $0.325$\\
MEs & $z_{l}$ & $0.427$ & $0.230$ & $0.144$ & $0.509$ & $0.242$ &
$0.020$ & $3.866$ & $0.483$\\
\hline
\end{tabular}\\
\medskip{Note: Quantities $z_{i}$, $z_{a}$, $z_{f}$ and $z_{l}$ are the
identity, assembly, formation and last merging times, respectively.\\
MEs$^{*}$ refers to a sample where only haloes more massive than $5 \times
10^{12}h^{-1}$M$_{\odot}$ have been considered.}
\end{minipage}
\end{table*}
\subsection{Formation and Evolution Times}
\citet{DeLucia:2007p414} defined a set of times related to formation and
evolution of dark matter haloes and galaxies that they reside in. We adopt the
same definitions for convenient and easy comparison, but we add one more quantity
named last merging time, $z_{l}$. Briefly, the different times are defined as
follows:
\begin{itemize}
\item Assembly time ($z_{a}$) is the redshift when $50$ per cent of the final
stellar mass is already present in a single galaxy of the merger tree.
\item Identity time ($z_{i}$) is the redshift when the last major (the two
progenitors both contain at least $20$ per cent of the stellar mass of the
descendant galaxy) merger occurred.
\item Formation time ($z_{f}$) is the redshift when $50$ per cent of the mass
of the stars in the final galaxy at $z = 0$ have already formed.
\item Last merging time ($z_{l}$) is the redshift when the last merger
occurred.
\end{itemize}
We compute all four times, measured in redshifts, related to the formation and
evolution of galaxies, and present numerical results in Table \ref{tb:times}.
Fig. \ref{f:formationetcTimes} shows cumulative distributions of the assembly
($z_{a}$), identity ($z_{i}$), formation ($z_{f}$), and last merging ($z_{l}$)
times. In Table \ref{tb:timeKSresults} we show probabilities that the
distributions of formation and evolution times of IfEs and MEs are drawn from the
same parent distribution. The KS tests indicate great differences in all cases,
implying that the formation and evolution of IfEs is different from MEs.
\begin{table}
\caption{Results of the Kolmogorov$-$Smirnov (KS) test: isolated field
ellipticals (IfEs) vs. the control sample (MEs) galaxies.}
\label{tb:timeKSresults}
\begin{tabular}{ccr}
\hline
Time & D$-$value & KS Probability (p$-$value)\\
\hline
$z_{a}$ & $0.190$ & $5.7 \times 10^{-8}$\\
$z_{i}$ & $0.239$ & $1.1 \times 10^{-5}$\\
$z_{f}$ & $0.353$ & $< 10^{-10}$\\
$z_{l}$ & $0.179$ & $6.5 \times 10^{-8}$\\
\hline
\end{tabular}\\
\medskip{Note: Quantities $z_{i}$, $z_{a}$, $z_{f}$ and $z_{l}$ are the
identity, assembly, formation and last merging times, respectively.}
\end{table}
Fig. \ref{f:formationetcTimes} (top panel) shows that isolated field elliptical
galaxies have assembled at lower redshifts than galaxies of MEs. Thus, in
general, stars of IfEs form later than stars of MEs and thus IfEs are younger.
This result is natural for hierarchical Cold Dark Matter models. If the
conventional theory, that higher density areas collapse earlier, holds, this
implies that IfEs have formed originally in less dense regions than MEs. Fig.
\ref{f:formationetcTimes} (second panel) also shows that IfEs undergo their major
merging events at significantly lower redshifts than the ellipticals of MEs. The
difference in redshift is significant (see Table \ref{tb:times}) and implies that
a different formation mechanism is behind the evolution of IfEs and MEs.
Fig. \ref{f:formationetcTimes} (third panel) shows that stars that will
eventually form an IfE galaxy are present already at higher redshifts than for
MEs, in agreement with observational findings \cite[e.g.][]{Reda:2005p561}. These
results indicate that IfEs can form stars more efficiently than MEs (see also
Fig. \ref{f:massAssembly}). Note however, that we find the formation time
($z_{f}$) to be highly dependent on the mass of the dark matter halo, as noted in
\citet[][]{DeLucia:2006p376}. If we limit MEs to galaxies with dark matter halo
mass greater than $5 \times 10^{12}h^{-1}$M$_{\odot}$ we find the median
formation time to be at very high redshift ($z \sim 3.3$). This shows that
galaxies in massive dark matter haloes form the bulk of their stars already at
very early cosmic epochs, in agreement with the general hierarchical halo mass
growth scenario.
\begin{figure}
\includegraphics[width=84mm]{times.eps}
\caption{Cumulative distributions of assembly ($z_{a}$), identity ($z_{i}$),
formation ($z_{f}$) and last merging ($z_{l}$) times measured in redshifts. The
blue solid lines corresponds to IfEs, while the red dashed lines mark the MEs.}
\label{f:formationetcTimes}
\end{figure}
We define a major merger as an event where the two progenitors both contain at
least $20$ per cent of the stellar mass of the descendant galaxy. Almost half
($\sim 46.3$ per cent) of isolated field elliptical galaxies have experienced a
major merger at some point of their formation history. The percentage of major
merging events for IfEs is higher than for MEs, where only one third ($\sim 33.3$
per cent) of galaxies have experienced a major merging event. If all non-isolated
E-type galaxies, independent of their luminosity or status in their dark matter
haloes, are considered, only about four per cent experience a major merger. The
difference is significant and shows that it is possible to form elliptical
galaxies, isolated or not, without a major merger via disc instabilities.
Let us study the redshifts of the last merging event $z_{l}$ (see Fig.
\ref{f:formationetcTimes} bottom panel and Tables \ref{tb:times} and
\ref{tb:timeKSresults}) and the total number of merging events. IfEs have their
last merging event at lower redshifts than MEs. The median redshift of the last
merging event is $0.21$ and $0.28$ for IfEs and MEs, respectively. This shows
that IfEs have merging activity also at very late stages of their evolution in
agreement with \cite{HernandezToledo:2008p648}. \cite{Hau:1999p629} quoted an age
estimate of $\sim 1.1$ Gyr for isolated elliptical galaxies since the last
merger. This age estimate from observational data seems to disagree with our
median time of the last merging, which is more than twice as high: $\sim 2.5$
Gyr. However, we do find IfEs that have experienced their last merging event only
$\sim 0.3$ Gyr ago. Thus, the discrepancy can arise from the small number of
observed IfEs in \cite{Hau:1999p629}.
It is possible that evolution of IfEs is suppressed by the low density
environment they reside in, thus the major merging events happen later stage in
their evolution if at all. The larger fraction of IfEs having major mergers than
MEs can also be due to the requirement of the magnitude gap not only the region
they formed in. IfEs of denser areas with comparable sized galaxies must clean
their environment effectively, leading to a larger fraction of major merging
events.
While studying the redshifts of the last merging events we noticed that $6$ IfEs
did not experience a single merging event during their evolution. These galaxies
have developed in truly isolated areas; although, maybe, a greater mass
resolution in simulations would reveal one or more minor merging events. Despite
the resolution limitations, the result is in agreement with observations where
some IfEs do not show any signs of merging activity. Typically, simulated IfEs
experience one to $10$ merging events (above the mass limit of
$10^{8}h^{-1}$M$_{\odot}$) during their lifetime, the average being $7.6$ and the
median six mergers. These numbers are significantly higher than for MEs, for
which we find the average and median of $4.6$ and one merging event,
respectively. This suggests that IfEs accrete multiple smaller haloes (and
galaxies) during their formation, cleaning their neighbourhood from dwarf
galaxies effectively. This leads to a question how effectively IfEs accrete mass
as a function of time compared to MEs.
\subsection{Mass Assembly}
Fig. \ref{f:massAssembly} shows the dark and stellar matter mass assembly as a
function of redshift for both IfEs and MEs. The figure shows that IfEs start to
form around the same epoch as MEs, however, IfEs accumulate stellar matter much
faster. According to our findings IfEs can form stars more efficiently than MEs,
while MEs seem to accrete dark matter slightly faster than IfEs. We confirm that
IfEs form the bulk of their stars at $z > 2$, as suggested in
\cite{Reda:2005p561}. We also note that at redshift one IfEs have formed over
half of their stars (stellar mass) and have gathered as much as $80$ per cent of
their final dark matter. The galaxies of MEs have accreted roughly the same
fraction of dark matter as IfEs at $z \sim 1$, however, they have formed as
little as $30$ per cent of their stars compared to the final stellar matter at $z
= 0$. This difference is significant and shows that IfEs form their stars quickly
and are rather dark-matter poor compared to other elliptical galaxies (see also
Fig. \ref{f:massComparison} for stellar vs. dark matter). It is also noteworthy
that IfEs continue to accrete dark matter till $z = 0$ while MEs have gathered
$\sim 99$ per cent of their final dark matter already at $z \sim 0.5$. All
this points towards a different formation mechanism for these two galaxy classes and
suggests that late merging events can be a significant part of the evolution of
an IfE.
\begin{figure}
\includegraphics[width=84mm]{massAssembly.eps}
\caption{Mass assembly of isolated field ellipticals (IfEs) and MEs as a function
of redshift $z$. Lines show the median value at given redshift.}
\label{f:massAssembly}
\end{figure}
To further illustrate the mass assembly of IfEs, we plot the sample of blue,
light and faint IfEs separately in Fig. \ref{f:binnedMassAssembly}. We have
divided the IfEs sample into two: light ($M_{DM} \leq 10^{12}h^{-1}$M$_{\odot}$)
and heavy ($M_{DM} > 10^{12}h^{-1}$M$_{\odot}$) IfEs. Fig
\ref{f:binnedMassAssembly} shows that there is a significant difference in mass
accretion of light and heavy IfEs. The stellar mass of heavy IfEs follows closely
the dark matter mass accretion, while light IfEs evolve differently. The heavy
IfEs have created half of their stellar matter already at $z \sim 1.6$, while the
light IfEs have created barely $10$ per cent of their stellar matter, in
agreement with the findings of \cite{Treu:2005p660}. We note that the light IfEs
create half of their stellar matter by $z \sim 0.7$. The heavy IfEs also form
stars extremely efficiently compared to the light IfEs, as the stellar mass
follows dark matter accretion closely. Fig. \ref{f:binnedMassAssembly} also shows
that heavy IfEs host older stellar populations than light IfEs, as $\sim 99$ per
cent of their stellar mass has been accumulated already by $z \sim 0.3$. Thus,
more massive (and luminous) IfEs have old stellar populations, while the lighter
ones have formed a signification fraction of their stellar mass relatively
recently.
\begin{figure}
\includegraphics[width=84mm]{binnedMassAssembly.eps}
\caption{Mass assembly of isolated field ellipticals (IfEs) as a function of
redshift $z$ when the sample of IfEs has been divided by dark matter halo mass.
Lines show the median value at a given redshift.}
\label{f:binnedMassAssembly}
\end{figure}
\subsection{Merger Histories of IfEs}\label{s:mergerHistories}
So far we have shown that minor and major merging events take place during the
formation of an IfE. Thus, in the following subsections, we try to identify
general formation mechanisms based on different types of merger trees and merging
events. We study merger trees of all IfEs individually and identify three typical
formation scenarios.
We illustrate merger histories in merger plots. The isolated field elliptical
itself lies at the bottom of the plot at $z = 0$, and all its progenitors (more
massive than $10^{8}h^{-1}$M$_{\odot}$) are plotted upwards going back in time.
Galaxies with stellar mass larger than $10^{9}h^{-1}$M$_{\odot}$ are shown as
symbols, and are colour-coded as a function of their rest-frame $B - R$ colour.
\subsubsection[]{Solitude}\label{Form1}
Fig. \ref{f:alone} shows an example of an isolated field elliptical galaxy that
has developed undisturbed, alone, and has not undergone even a single merging
event. We group IfEs that form in this fashion to a single group that we call
solitude formation class.
Even though the Fig. \ref{f:alone} does not show a single merging event during
the formation history, this may not be the whole story; the plot does not show
mergers smaller than the resolution limit ($\sim 10^{8}h^{-1}$M$_{\odot}$) of the
Millennium Simulation. Therefore, it is possible or even likely that IfEs
belonging to this class have had one or even several minor merging events during
their evolution. Even so, the merging events would have involved very light dark
matter clumps, and it is likely that no significant observational evidence would
exist. In this context small merging events can be interpreted as accretion of
dark matter making solitude a proper formation mechanism.
We identify six IfEs, corresponding to mere two per cent of all IfEs, that belong
to the solitude class. This implies that either IfEs that develop truly alone in
underdense regions are extremely rare or that some of the IfEs have been
misclassified. It is possible that some IfEs with one or more small mergers could
belong to this formation class, especially if the motivation of classification is
based on observability of each class. It is not obvious how massive a merging
event is required to find observational evidence of a merger complicating the
classification.
The six IfEs identified all show unsteady evolution in their colour as a function
of redshift. This is somewhat surprising as one would assume that a passively
evolving galaxy should show a steady colour evolution from blue to red while the
stellar population ages. The bulge formation of solitude class is assumed to
happen via disc instabilities. Solitude IfEs reside inside lighter dark matter
haloes than IfEs of other class, with typical dark matter halo mass of $ \sim 2
\times 10^{11}h^{-1}$M$_{\odot}$ or even less. This makes IfEs that belong to the
solitude class the lightest IfEs in the MS. The formation, when the dark matter
mass grows larger than $10^{8}h^{-1}$M$_{\odot}$, epoch of solitary IfEs is in
the redshift range $3.6 < z < 5.3$. However, we find one solitude IfE that formed
as late as $z \sim 2.5$, making it a very late bloomer, showing that, IfEs can
have formed very recently. Solitude IfEs are in agreement with observations of
e.g. \citet[][]{Aars:2001p564} and \citet[][]{Denicolo:2005p568}, who did not
find any evidence of merging activity while studying their samples of IfEs.
\begin{figure}
\includegraphics[width=84mm]{aloneEDIT.ps}
\caption{Example of a merger tree of an isolated field elliptical galaxy that
has developed alone without any significant merging events. Symbols are
colour-coded as a function of the $B - R$ colour and their area scales with the
stellar mass. Only progenitors more massive than $10^{8}h^{-1}$M$_{\odot}$
are shown.}
\label{f:alone}
\end{figure}
\subsubsection[]{Coupling}\label{Form2}
Fig. \ref{f:equal} shows an example of an isolated field elliptical that has
undergone at least one ``equal" size merger during its evolution. IfEs that have
experienced at least one equal sized merger comprise a formation class named
coupling.
Our definition for an equal size merging is the following: the lighter of the
merged galaxies had at least $50$ per cent of the stellar mass of the more
massive one. This is far larger than in our definition of a major merger ($20$
per cent) that was applied when calculating the identity time $z_{i}$. Thus, an
equal sized merger guarantees that the merging event has had a great impact not
only on the morphology, but also on the evolution and other properties of the
IfE.
The motivation for this formation class resides in observations; a major merging
event should be visible in observational data. Thus, an equal sized merging
should leave distinct marks to the descendant galaxy, which should be observable
for at least few Gyrs if not longer. It has been suggested that isolated field
ellipticals have formed in relatively recent mergers of spiral galaxy pairs, i.e.
in merging of two comparable sized galaxies. \cite{Marcum:2004p572},
\cite{Reda:2007p559} and \cite{Kautsch:2008p615} have found several isolated
galaxies that show signs of recent morphological disturbances, while
\cite{Hau:1999p629} \citep[see also][]{Hau:2006p557} have found that $\sim 40$
per cent of isolated galaxies show kinematically distinct cores (KDC). KDCs are
generally believed to be the result of a major or an equal sized merger
\citep{Hernquist:1991p641}. Thus, it is possible that these observations have
already identified galaxies that belong to this formation class. However, it is
also possible that KDCs can form in an early collapse without subsequent mergers
\citep{Harsoula:1998p631}, complicating the identification of the formation
mechanism.
We identify in total $93$ IfEs, corresponding to $\sim 32$ per cent of all IfEs,
that belong to the coupling class of formation scenarios. IfEs that belong to the
coupling formation class show colour evolution that is in agreement with
conventional theory. The time from the last equal sized merging correlates well
with the colour of the galaxy at $z = 0$; galaxies with late merging are bluer
than galaxies that experienced their equal sized merging a long time ago. The
mean redshift of the last equal sized merger is $1.09$ while the median is
$0.62$. IfEs that belong to this formation class start to form usually in the
redshift range $3.6 < z < 5.3$. We also identify a few IfEs that have started to
form as early as $z \sim 8$, making IfEs that belong to the coupling class older,
in a statistical sense, than IfEs that belong to the solitude formation class.
The coupling formation mechanism can explain several observed IfEs. The
colour-magnitude diagrams of \cite{Reda:2004p502} show a slope and scatter that
is in agreement with equal-mass mergers. They find $11$ per cent of their IfEs to
show boxy isophotes that can form when equal sized galaxies merge.
\cite{Marcum:2004p572} argue that all except one of their IfE have luminosities
that would, at most, be consistent with a single equal-mass merger event. Note
that all these observations are best explained by the IfEs of the coupling class.
\begin{figure}
\includegraphics[width=84mm]{equalEDIT.ps}
\caption{Example of a merger tree of an isolated field elliptical galaxy that has
undergone an equal sized merger. Symbols are colour-coded as a function of the
$B - R$ colour and their area scales with the stellar mass. Only progenitors more
massive than $10^{8}h^{-1}$M$_{\odot}$ are shown.}
\label{f:equal}
\end{figure}
\subsubsection[]{Cannibalism}\label{Form3}
Fig. \ref{f:cannibal} shows a typical isolated field elliptical galaxy that has
developed and accreted dark matter and stellar mass through multiple, small and
larger merging events, but has not experienced any equal sized mergers. All IfEs
that form via multiple mergers form a formation class called cannibalism.
The merger trees of cannibal IfEs show a large number of mergers, with also a
relatively large ones. It is obvious from the merger trees that the colour of the
IfE does not change due to a minor merger. This is due to the fact that many
small merging events are likely to be dry, and therefore, do not induce
significant star formation that would make the global colour of the IfE bluer.
However, the largest merging events can have a significant impact on the IfE's
morphology and colour (see Fig. \ref{f:cannibal} and the merging event around 0.8
Gyr ago). Some merging events, visible in the merger trees of cannibal IfEs, can
be classified as major mergers, and these could be visible in observational data.
Most ($194$) of the IfEs in the MS belong to the cannibal class, corresponding to
$\sim 66$ per cent of all IfEs. The large number of IfEs belonging to this class
shows that merging events are important for the formation and evolution of
isolated field elliptical galaxies. Cannibals often show a large number of
merging events, more than $20$, while the number of major merging events ranges
from one to three. The possibility of major mergers can complicate the
identification of IfEs of this class, especially as we cannot identify any
preferred time for the last major merger. In general, IfEs of the cannibal class
form early, in the redshift range $5 < z < 12$. However, we identify few cannibal
IfEs that form as late as $z \sim 2.4$. The cannibal IfEs that form late could
also be classified as solitudes, especially as these IfEs do not, in general,
show a single major merging event only few minor ones.
This formation class can explain a few observed IfEs. \cite{Reda:2004p502} found
low-luminosity dwarfs close to an isolated galaxy that avoid accretion. Three of
their IfEs show high central space density, but they also found quite strong
morphological disturbances requiring an accretion of a fairly large galaxy. This
is in agreement with cannibal IfEs that have had a major merging event that would
explain the morphological disturbances, while a large number of smaller satellite
dwarfs could have survived due to the large dark matter halo and long dynamical
friction times.
\begin{figure}
\includegraphics[width=84mm]{cannibalEDIT.ps}
\caption{Example of a merger tree of an isolated field elliptical galaxy that has
undergone multiple merging events, but not any equal sized ones. Symbols are
colour-coded as a function of the $B - R$ colour and their area scales with the
stellar mass. Only progenitors more massive than $10^{8}h^{-1}$M$_{\odot}$ are
shown.}
\label{f:cannibal}
\end{figure}
\section[]{Discussion}\label{s:discussion}
\subsection{Population of Blue and Faint IfEs}
Figs. \ref{colourMag} and \ref{colourMag2} show a separate population of blue
and faint IfEs. This population was also confirmed to comprise isolated field
elliptical galaxies that reside in light, $M_{DM} < 10^{12}h^{-1}M_{\odot}$,
dark matter haloes. Moreover, the IfE samples of \citet{Reda:2004p502} and
\citet{Marcum:2004p572} do not reveal a single observed IfE that would belong to
this predicited population of IfEs.
\citet{HernandezToledo:2008p648} studied isolated galaxies and used modern
observations of Sloan Digital Sky Survey (SDSS) Data Release (DR) 6. The base
sample of their galaxies were the Catalogue of Isolated Galaxies (CIG) in the
Northern Hemisphere combiled by \citet{1973AISAO...8....3K}. Although the
isolation criteria in the study of \citet{1973AISAO...8....3K} and in our study
are not the same, it is still interesting to compare if any of the elliptical
galaxies could belong to the population of faint and blue IfEs that our results
predict. \citet{HernandezToledo:2008p648} noted that four early-type galaxies of
the CIG sample showed blue ($g - i < 1.0$) colours. Three of their blue galaxies
were found to be disky while the other one was found to be boxy. From the
morphological inspections they concluded that three of the blue galaxies were
ellipticals while one was S$0$-type. Thus, below we will inspect if any of the
three blue elliptical galaxies noted in \citet{HernandezToledo:2008p648}, could
belong to the faint and blue population of IfEs that simulations predict.
As the magnitudes of \citet{HernandezToledo:2008p648} are in the SDSS system, we
need a method to transform their magnitudes for a comparison. For a crude
comparison we can use galaxy colours of \citet{Fukugita:1995p664} and derive a
transformations:
\begin{equation}\label{eq:r-mag}
M_R \approx M_r - 0.35
\end{equation}
for $R-$band magnitudes for a typical elliptical. Because of our IfEs show bluer
colours than typical ellipticals, this transformation is not very accurate,
however, for our purposes it should be adequate. To transform the SDSS colours,
we use the following equation:
\begin{equation}\label{eq:colour}
B - R \approx (g - i) + 0.44 \quad ,
\end{equation}
which has been derived from the works of \citet{1996AJ....111.1748F} and
\citet{Jester:2005p662}. Again, we note that this may not provide exact colour
transformation, but should provide satisfactory conversion for our crude
comparisons.
The three isolated ellipticals of \citet{HernandezToledo:2008p648} show $g - i$
colours of $0.91$, $0.98$ and $1.02$. With the equation \ref{eq:colour} we can
approximate that these colours correspond to $B - R$ colours of $1.35$, $1.42$,
and $1.46$. Clearly at least one of their galaxy is blue enough ($B - R \leq
1.4$) to belong to the separate population of faint and blue IfEs. Given the
inaccuracy of equation \ref{eq:colour} it is possible that all three of their
galaxies would belong to the blue population of IfEs. The blue galaxies of
\citet{HernandezToledo:2008p648} have absolute $r-$band magnitudes $-20.84$,
$-21.24$, and $-16.73$, respectively. With the help of equation \ref{eq:r-mag} we
can approximate their $R-$band magnitudes to be $-21.19$, $-21.59$, and $-17.08$,
respectively. Thus, at least two of their galaxies are faint enough ($M_{R} >
-21.5$) to be part of the population.
Above magnitudes and colours show that possibly at least one and up to three
isolated field elliptical galaxies that belong to the predicited population of
faint and blue IfEs has already been observed. One of the isolated ellipticals of
\citet{HernandezToledo:2008p648} is significantly fainter than any of our IfEs,
and would not qualify as an IfE in our study. Moreover, as the isolation criteria
of CIG galaxies are different that ours, we cannot conclude without a doubt that
any of the blue and faint population IfEs have been observed. Thus, more
observations of blue and faint isolated elliptical galaxies are required to
confirm our prediction.
\subsection{Observing a Formation Class}
Section \ref{s:mergerHistories} introduced three typical formation classes of
isolated field elliptical galaxies; solitude, coupling, and cannibalism. Our
results show that the majority of IfEs experience numerous merging events during
their evolution and belong to the cannibalism class. A smaller, yet a significant
fraction of IfEs belong to the coupling class, whose members have experienced an
equal sized merging, while only a small fraction of IfEs show insignificant
merging activity and belong to the solitude class. If merging events are so
important for the evolution of IfEs, an obvious question remains: Is it possible
to identify the formation scenario based on observational data? Therefore, in
this section we briefly discuss the properties of IfEs that can be observed and
that at the same time would readily indicate the formation mechanism of the IfE.
IfEs that belong to the coupling class form later than cannibal IfEs and are
found to reside in lighter dark matter haloes than cannibal IfEs: the median
masses are $54$ and $200 \times 10^{10}h^{-1}$M$_{\odot}$, respectively. The
coupling class IfEs can therefore be merger remnants of nearby galaxy pairs
having a modest sized dark matter halo at redshift zero. Unfortunately, dark
matter haloes cannot be observed directly. IfEs of equal sized mergings can show
a weak X-ray emission; however, it is not clear how well this could be detected
if at all. Moreover, differentiating coupled IfEs from cannibals might be
difficult from X-ray data only.
The mean number of companion galaxies of IfEs that have formed through coupling
is lower than for the cannibal IfEs, but larger than for the solitary IfEs.
Unfortunately, the distribution of stellar mass, $B-$band magnitude and colour of
IfEs of the coupling class is indistinguishable from the cannibal IfEs. The
redshifts of equal sized merging events suggest that to be able to identify
galaxies belonging to the coupling class, observations should concentrate on
intermediate redshifts in the range $0.25 < z < 1.4$. However, to find evidence
of equal sized merging may not be simple; it is not clear how different are the
traces an equal size merger leaves, compared to a major merger. Study of stellar
populations could provide a way to identify IfEs that have experienced an equal
sized merging, but this may not be applicable for high redshifts.
A cannibal IfE can be a merger remnant of a small or compact group. We find
cannibal IfEs to reside in more massive and larger dark matter haloes than other
types of IfEs. However, their dark matter haloes are less massive than haloes of typical
groups, leaving only small or compact groups to consider. Multiple merging events
of cannibal IfEs could also be visible in their stellar populations, especially
if merging events were gas rich, so-called wet mergers. Observational evidence
shows that it is possible to detect IfEs in X-ray observations
\citep[e.g.][]{Mulchaey:1999p563, Sivakoff:2004p628, OSullivan:2004p569,
Memola:2009p642}. Due to their more massive dark matter haloes, X-ray bright IfEs
are the best candidates for the cannibal formation class.
\subsection[]{Comparison to Fossil Groups}
Have fossil groups formed in a similar way as isolated field elliptical galaxies?
Are IfEs an intermediate product while they develop to become fossil groups, or
vice versa? These questions are justified as the selection criteria of IfEs are
very similar to those used for the identification of fossil groups, especially if
only optical data are available. Objects of both classes are required to show a
large magnitude gap between the brightest and the second brightest galaxy in
optical. In general, fossil groups are also required to show extended X-ray
emission while it is not demanded for IfEs. However, several IfEs have shown some
amount of extended X-ray emission even it is not required. Due to similar
selection criteria objects of both classes might have similar evolution and
assembly histories and a similar formation mechanism, being actually the same
class seen only at different phases of evolution.
The majority of fossil groups seem to have experienced the last major merging
event longer than $3$ Gyr ago and they have assembled half of their final mass by
$z \geq 0.8$ \citep{vonBendaBeckmann:2008p598}. Moreover,
\cite{vonBendaBeckmann:2008p598} found from simulations that only $15$ per cent
of fossil groups experience the last major merger less than $2$ Gyr ago, and at
least $50$ per cent had the last major merger longer than $6$ Gyr ago. These
timescales are in modest agreement with our findings. Our results show that IfEs
have experienced their last major merging, on average, $\sim 6$ Gyr ago, while
IfEs have half of their final mass assembled by $z \sim 1$. For fossil groups
this number is $\sim 0.6$ \citep[][]{DiazGimenez:2008p617}.
\cite{DiazGimenez:2008p617} study the evolution of the brightest galaxies of
fossil groups. They identified fossil groups from the Millennium Simulation and
adopted the same galaxy catalogue \citep{DeLucia:2007p414} as in this study,
making it interesting to compare their findings to ours.
When comparing the medians of assembly, formation and identity times we find that
isolated field elliptical galaxies have assembled their stars earlier than the
brightest galaxies of fossil groups; $z_{a} \sim 1.1$ and $\sim 0.6$,
respectively. However, fossil groups form significantly earlier than IfEs; $z_{f}
\sim 3.6$ and $\sim 1.5$, respectively. The median time of the last major merging
of the IfEs is almost twice ($z_{i} \sim 0.6$) the identity time of the brightest
galaxies of fossil groups ($z_{i} \sim 0.3$). These discrepancies in formation
and evolutionary times cast a serious doubt over the idea of similar formation
mechanisms.
\cite{vonBendaBeckmann:2008p598} argue that the primary driver for the large
magnitude gap is the early infall of massive satellites that is related to the
early formation time of fossil groups. We do not find infall of massive
satellites at early time in our sample of simulated isolated field elliptical
galaxies. The early evolution of an IfE includes relatively minor mergers, except
for some IfEs that experience an equal sized merging. However, the time of the
equal sized mergers is usually in a later stage of the development of the IfE,
not early as argued for fossil groups. Our study shows that the primary driver
for the large magnitude gap of IfEs is either a merging of a comparable sized
galaxy pair or effective mass accretion and sweeping up of surrounding galaxies
(coupling and cannibalism, respectively).
\citet{Dariush:2007p265} argue that fossil groups can be identified from dark
matter only simulations if one selects dark matter haloes more massive than $5
\times 10^{13}h^{-1}$M$_{\odot}$. They argue that above that limit, all optical
fossil groups in the Millennium Simulation have enough hot gas to show X-ray
emission, therefore qualifying as X-ray fossils. Our Table \ref{tb:properties}
shows that the median mass of the dark matter haloes of IfEs is only $\sim 1.2
\times 10^{12}h^{-1}$M$_{\odot}$, while even the $3$rd quartile is just $\sim 2.4
\times 10^{12}h^{-1}$M$_{\odot}$. This comparison shows that IfEs reside in
significantly lighter dark matter haloes than the brightest galaxies of fossil
groups. \citet{Dariush:2007p265} argued that fossil groups have formed early and
that more than $\sim 80$ per cent of their mass accumulated as early as $4$ Gyr
ago. Moreover, they suggest that X-ray fossil groups are not a distinct class of
objects but rather that they are extreme examples of groups which collapse early
and experience little recent growth. This formation scenario does not apply for
IfEs, as cannibal IfEs can have merging activity till the redshift $z \sim 0$,
and IfEs start to form later than fossil groups.
Although a large magnitude gap seen in both fossils and IfEs could imply a
similar formation mechanism, the above comparison does not support this idea.
Considering the assembly and formation times it seems unlikely that fossil
groups, and especially their brightest galaxies, could share the same formation
mechanism as isolated field elliptical galaxies. The large differences in dark
matter halo masses of fossil groups and the most massive IfEs makes it improbable
that fossil groups and IfEs are the same class of objects seen at different
phases of their evolution. We therefore conclude that fossil groups and IfEs form
two distinct classes.
\section[]{Summary and Conclusions}\label{summary}
The aim of this paper is twofold: 1) to compare simulated field elliptical
galaxies with observed ones and to make predictions on their properties, and 2)
to define the formation mechanism and history of isolated field elliptical
galaxies. We also discuss if isolated field elliptical galaxies are related to
fossil groups and if they can share a common formation mechanism.
Our results show that it is possible to identify isolated field elliptical
galaxies (IfEs) from cosmological $N-$body simulations with semi-analytical
models of galaxy formation, that have similar properties to observed IfEs. We
show that simulated IfEs are in good agreement with observations when similar
identification criteria are adopted. The colour-magnitude diagram of simulated
IfEs agree with observations, and the average colour of simulated and observed
IfEs are within their error limits when all simulated IfEs are considered.
Unfortunately, observational data sets are still small complicating more detailed
comparison. An agreement in age and mass estimates of IfEs between observations
and simulations is found. However, the age comparisons are less robust due to
different age definitions.
Our results show that isolated field ellipticals are very rare objects; we
find their total number density to be as low as $\sim 8.0 \times 10^{-6}$h$^{3}$
Mpc$^{-3}$. Our result agrees with observational estimates of number of IfEs,
which however, are inaccurate at best. Our IfEs have a small number of companion
galaxies, ranging from only a few dwarf companions to as much as about $20$.
Thus, IfEs are not completely isolated, although they are found to be located in
underdense regions. Our results show that IfEs reside in relatively light dark
matter haloes. However, at the same time, the baryonic to dark matter ratio is
higher in IfEs than in Es generally. The stellar mass of IfEs grows almost
linearly with the dark matter mass with a mass ratio (dark/stellar) of $\sim 4
\times 10^{-2}$, whereas our comparison ellipticals have a lower stellar to dark
matter ratio. Therefore, IfEs are good candidates for galaxies with low dark
matter mass to stellar light ratio. We also find a flat $B-$band luminosity
function for our simulated IfEs.
When studying the basic properties of IfEs we find that IfEs populate different
regions in colour-magnitude diagrams than regular elliptical galaxies, which are
found mostly in the red sequence. This is due to the bimodality of the
distribution of colours, as IfEs populate not only the red sequence, but also the
blue cloud. From simulation data we find a separate population of blue and
faint IfEs. On average IfEs are found to be bluer than our control sample
ellipticals, howevever, this result is biased because of the separate population
of IfEs. The bluest IfEs are found to reside in light dark matter haloes, while
red IfEs are usually found inside more massive ($M_{DM} \geq
10^{12}h^{-1}M_{\odot}$) dark matter haloes. We note that simulations predict a
previously unobserved population of blue, dim and light galaxies that fulfill
observational criteria to be classified as isolated field elliptical galaxies.
These galaxies have only a few companions, which are usually located many times
further away than the virial radius of their dark matter halo. These blue, dim
and light IfEs have formed their stars only lately and have continued to accrete
dark matter mass till redshift zero. This population of IfEs is very interesting
as it has not been detected yet in observations.
Our results also show that IfEs start to form around the same epoch as the
galaxies of the control sample (MEs); however, IfEs seem to accumulate stellar
mass much faster. IfEs form stars more efficiently than MEs, while MEs accrete
dark matter slightly faster than IfEs. We confirm that IfEs form the bulk of
their stars at $z > 2$, as suggested in \cite{Reda:2005p561}. By the redshift one
IfEs have formed over half of their stars (stellar mass) and have gathered as
much as $80$ per cent of their final dark matter. The galaxies of MEs sample have
accreted roughly the same fraction of dark matter as IfEs by $z \sim 1$. However,
they have formed as little as $30$ per cent of their stars compared to the final
stellar matter at $z = 0$. This difference shows that IfEs form their stars
quickly and are rather dark-matter poor compared to other elliptical galaxies.
Moreover, IfEs continue to accrete dark matter till $z = 0$ while MEs have
gathered $\sim 99$ per cent of their final dark matter already by $z \sim 0.5$.
We note that more massive (and luminous) IfEs have older stellar populations,
while the lighter ones have formed a signification fraction of their stellar mass
relatively recently.
While studying the evolution of IfEs we note that almost half ($\sim 46$ per
cent) of IfEs have experienced at least one major merger during their formation
history, while only about four per cent (Es) up to one third (MEs) of control
sample ellipticals experience a major merger. Major merging events happen later
in IfEs evolution than for control sample galaxies, the average of the latest
major merging being $z \sim 0.6$ for IfEs while it is $z \sim 1.1$ for MEs. We
also find IfEs that have not experienced a single merger event above the mass
resolution limit of the MS during their evolution. Therefore, it is possible to
form elliptical galaxies without major mergers.
When inspecting the merger trees of simulated IfEs, we identify three typical
formation scenarios: solitude, coupling, and cannibalism, which can all lead to a
formation of an IfE. The scenarios range from a solitary growth (solitude class)
with quiet mass accretion and star formation to more violent evolution with
multiple mergers (cannibalism class). We also identify a formation scenario where
two comparable sized galaxies merge to form an IfE (coupling class). Our merger
trees show that merging events are important for IfEs that form through
cannibalism or coupling. All three formation classes are in agreement with
observational findings.
Comparison between isolated field elliptical galaxies and fossil groups show that
these two classes are distinct. Galaxies of both classes show some similarities,
but many properties and evolution times are significantly different. IfEs reside
in significantly lighter dark matter haloes and we do not find an infall of
massive galaxies at early times for IfEs as argued for fossil groups
\citep{vonBendaBeckmann:2008p598}. However, we cannot exclude that some fossil
groups could not share the formation mechanism of the most massive IfEs, namely
cannibalism.
\section*{Acknowledgements}
SMN acknowledges the funding by the Finnish Academy of Science and Letters and
the Nordic Optical Telescope Scientific Association (NOTSA). SMN would like to
thank Drs. Elena D'Onghia and Henry Ferguson for enlightening and inspiring
conversations, Dr. Gerard Lemson for invaluable help with the Millennium
Simulation database, and Ms Carolin Villforth for multiple inspiring
conversations. ES thanks the University of Valencia (Vicerrectorado de
Investigaci\'on) for a visiting professorship, the support by the Estonian
Science Foundation grant No. 8005 and by the Estonian Ministry for Education and
Science, grant SF0060067s08. We thank the anonymous referee for detailed reading
of the manuscript and comments that helped us to improve the original manuscript.
The Millennium Simulation databases used in this paper and the web application
providing online access were constructed as part of the activities of the German
Astrophysical Virtual Observatory.
|
\section{Introduction}
\label{sec:intro}
String theory, as a candidate theory of all fundamental interactions
including gravity, is obliged to contain patterns consistent with
observation. This includes the standard model gauge group and
particle content, as well as an extremely small cosmological
constant. If one also assumes that nature contains a low energy
supersymmetry, one would want to find patterns which are
\textit{qualitatively} close to the MSSM. An even more ambitious
goal would be to find a theory consistent with the standard model
spectrum of masses and a prediction for sparticle masses which can
be tested at the LHC. Some progress has been made to this end
starting from different directions~\cite{Cleaver:1998saa,
Braun:2005ux, Cvetic:2005bn, Buchmuller:2005jr, Buchmuller:2006ik,
Lebedev:2006kn, Kim:2007mt, Chen:2007px, Lebedev:2007hv,
Lebedev:2008un, Beasley:2008dc, Donagi:2008ca, Beasley:2008kw,
Donagi:2008kj, Blumenhagen:2008zz, Bourjaily:2009vf, Hayashi:2009ge,
Huh:2009nh, Marsano:2009gv}, i.e. free fermionic, orbifold or smooth
Calabi-Yau constructions of the heterotic string, intersecting
D-brane constructions in type II string, and M or F theory
constructions. Much of this progress has benefited from the
requirement of an intermediate grand unified gauge symmetry which
naturally delivers the standard model particle spectrum.
In this paper we focus on the ``mini-landscape" of heterotic
orbifold constructions~\cite{Buchmuller:2005jr, Buchmuller:2006ik,
Lebedev:2006kn, Lebedev:2007hv, Lebedev:2008un}, which give several
models which pass a significant number of phenomenological
hurdles.\footnote{For reviews, see
\cite{Nilles:2008gq,Raby:2009dm}.} These models have been analyzed
in the supersymmetric limit. They contain an MSSM spectrum with
three families of quarks and leptons, one or more pairs of Higgs
doublets and an exact R parity. In the orbifold limit, they also
contain a small number of vector-like exotics and extra $U(1)$ gauge
interactions felt by standard model particles. These theories also
contain a large number of standard model singlet fields, some of
which are moduli, i.e.\thinspace blow up modes of the orbifold fixed points.
The superpotential for these orbifold theories can be calculated
order by order in powers of products of superfields. This is a
laborious task which is simplified by assuming that any term allowed
by string selection rules appears with an order one coefficient in
the superpotential. With this caveat it was shown that all
vector-like exotics and additional $U(1)$ gauge bosons acquire mass at
scales of order the string scale at supersymmetric minima satisfying
$ F_I = D_a = 0 $ for all chiral fields labeled by the index $I$ and
all gauge groups labeled by the index $a$. In addition, the value
of the gauge couplings at the string scale and the effective Yukawa
couplings are determined by the presumed values of the vacuum
expectation values [VEVs] for moduli including the dilaton, $S$, the
bulk volume and complex structure moduli, $T_i, i = 1,2,3$ and $U$ and the
SM singlet fields containing the blow-up moduli
\cite{Nibbelink:2007pn,Nibbelink:2009sp}. Finally the theories also
contain a hidden sector non-Abelian gauge group with QCD-like chiral
matter. The problem which has yet to be addressed is the mechanism of
moduli stabilization and supersymmetry breaking in the
``mini-landscape" models.\footnote{For a preliminary analysis, see
\cite{Lebedev:2006tr}. Also moduli stabilization and supersymmetry
breaking in Type II string models and F theory constructions have
been considered in
\cite{Giddings:2001yu,Kachru:2003aw,Heckman:2008es,Marsano:2008jq,Heckman:2008qt}.}
In this paper we focus on the problem of moduli stabilization and
SUSY breaking in the context of heterotic orbifold models. In
Section \ref{sec:general_structure} we summarize the general
structure of the K\"{a}hler and superpotential in heterotic orbifold
models. The models have a perturbative superpotential satisfying
modular invariance constraints, an anomalous $U(1)_A$ gauge symmetry
with a dynamically generated Fayet-Illiopoulos $D$-term and a hidden
QCD-like non-Abelian gauge sector generating a non-perturbative
superpotential. In Section \ref{sec:one_gaugino_condensate} we
consider a simple model with a dilaton, $S$, one volume modulus,
$T$, and three standard model singlets. The model has only one
gaugino condensate, as is the case for the ``benchmark models" of
the ``mini-landscape" \cite{Lebedev:2007hv}. We obtain a `hybrid
KKLT' kind of superpotential that behaves like a single-condensate
for the dilaton $S$, but as a racetrack for the $T$ and, by
extension, also for the $U$ moduli; and an additional matter $F$
term, driven by the cancelation of an anomalous $U(1)_A$ $D$-term,
is the seed for successful up-lifting. Previous analyses in the
literature have also used an anomalous $U(1)_A$ $D$-term in
coordination with other perturbative or non-perturbative terms in
the superpotential to accomplish SUSY breaking and
up-lifting~\cite{Binetruy:1996uv,Dvali:1996rj,Dvali:1997sr,Lalak:1997ar,deCarlos:2005kh,Choi:2006bh,Dudas:2006vc,Lalak:2007qd,Dudas:2007nz,Gallego:2008sv,Dudas:2008qf}.
We save a brief comparison of our work with some of these former
analyses for Section \ref{sec:one_gaugino_condensate}. In Section
\ref{sec:moduli_stabilization} we discuss the other moduli and their
stabilization. We conclude that a single gaugino condensate is
sufficient to break supersymmetry, stabilize all the moduli and
generate a de Sitter vacuum. Finally in Section \ref{sec:spectrum}
we evaluate the SUSY particle spectrum relevant for the LHC. The
main results from this analysis are listed in Tables
\ref{tab:weak_obs}, \ref{tab:param_space}.
\section{General structure}
\label{sec:general_structure}
In this section we consider the supergravity limit of heterotic
orbifold models. However, we focus on the ``mini-landscape" models
for definiteness. We discuss the general structure of the K\"{a}hler
potential, $\ensuremath{\mathcal{K}}$, the superpotential, $\ensuremath{\mathcal{W}}$, and gauge kinetic
function, $f_a$ for generic heterotic orbifold models. The
``mini-landscape" models are defined in terms of a \ensuremath{\mathbb{Z}_6\text{-II }}
orbifold of the six internal dimensions of the ten dimensional
heterotic string. The orbifold is described by a three dimensional
``twist'' vector $v$, which acts on the compact directions. We
define the compact directions in terms of complex coordinates:
\begin{eqnarray} \nonumber
Z_1 &\equiv& X_4 + iX_{5}, \\
Z_2 &\equiv& X_6 + iX_{7}, \\ \nonumber
Z_3 &\equiv& X_8 + iX_{9}.
\end{eqnarray}
The twist is defined by the action $Z_i \rightarrow e^{2\pi i v_i}
Z_i$ for $i = 1,2,3$, and for \ensuremath{\mathbb{Z}_6\text{-II }} we have $v =
\frac{1}{6} ( 1, 2, -3 )$ or a (60$^\circ$, 120$^\circ$,
180$^\circ$) rotation about the first, second and third torus,
respectively. This defines the first twisted sector. The second
and fourth twisted sectors are defined by twist vectors $2 v$ and $4
v$, respectively. Note, the third torus is unaffected by this
twist. In addition, for the third twisted sector, generated by the
twist vector $3 v$, the second torus is unaffected. Finally the
fifth twisted sector, given by $5 v$ contains the $CP$ conjugate
states from the first twisted sector. Twisted sectors with
un-rotated tori contain $N = 2$ supersymmetric spectra. This has
consequences for the non-perturbative superpotential discussed in
Section \ref{sec:gaugekinetic}. Finally, these models have three
bulk volume moduli, $T_i, \ i=1,2,3$ and one bulk complex structure
modulus, $U$, for the third torus.
\subsection{Anomalous $U(1)_A$ and Fayet-Illiopoulos $D$-term}
\label{sec:anomalous}
The orbifold limit of the heterotic string has one anomalous
$U(1)_A$ symmetry. The dilaton superfield $S$, in fact, transforms
non-trivially under this symmetry. Let $V_A, \ V_a$ be the gauge
superfields with gauge covariant field strengths, $W^\alpha_A,
W^\alpha_a$, of gauge groups, $U(1)_A, \ {\cal G}_a$, respectively.
The Lagrangian in the global limit is given in terms of a K\"{a}hler
potential
\cite{Green:1984sg,Fayet:1974jb,Dine:1987xk,Atick:1987gy,Dine:1987gj}
\begin{equation} \ensuremath{\mathcal{K}} = - \log(S + {\overline S} - \delta_{GS} V_A)
+ \sum_{a} ({\overline Q}_a e^{V_a + 2 q_a V_A} Q_a + {\overline
{\tilde Q}}_a e^{- V_a + 2 \tilde q_a V_A} \tilde Q_a )
\label{eq:Kahler} \end{equation} and a gauge kinetic superpotential \begin{eqnarray} \ensuremath{\mathcal{W}} =
& \frac{1}{2} [ \frac{S}{4} ( \sum_a k_a Tr W^\alpha_a W_{\alpha a}
+ k_A Tr W^\alpha_A W_{\alpha A} ) + h.c. ] . & \end{eqnarray} Note $q_a, \
\tilde q_a$ are the $U(1)_A$ charges of the `quark', $Q_a$, and
`anti-quark', $\tilde Q_a$, supermultiplets transforming under
${\cal G}_a$.
Under a $U(1)_A$ super-gauge transformation with parameter
$\Lambda$, one has
\begin{eqnarray} \label{eq:u1A} \nonumber
\delta_A V_A &=& -i (\Lambda - \bar \Lambda)/2, \\
\delta_A S &=& -i \frac{\delta_{GS}}{2} \Lambda ,
\end{eqnarray}
and
\begin{equation}
\delta_A \Phi = i q_\Phi \Lambda \Phi
\end{equation}
for any charged multiplet $\Phi$. The
combination \begin{equation} S + {\overline S} - \delta_{GS} V_A \end{equation} is $U(1)_A$
invariant. $\delta_{GS}$ is the Green-Schwarz coefficient given by
\begin{equation} \delta_{GS} = 4 \frac{Tr Q_A}{192 \pi^2} =
\frac{(q_a + \tilde q_a) N_{f_a}}{4 \pi^2} \label{eq:GS}
\end{equation} where the middle term is for the
$U(1)_A-$gravity anomaly and the last term is for the $U(1)_A \times
({\cal G}_a)^2$ mixed anomaly.
The existence of an anomalous $U(1)_A$ has several interesting
consequences. Due to the form of the K\"{a}hler potential (Eqn.
(\ref{eq:Kahler})) we obtain a Fayet-Illiopoulos $D$-term given by \begin{equation}
\xi_A = \frac{\delta_{GS}}{2(S + {\overline S})} = - \frac{1}{2}
\delta_{GS} \
\partial_S \ensuremath{\mathcal{K}} \end{equation} with the $D$-term contribution to the scalar potential given by
\begin{equation} V_D = \frac{1}{S + {\overline S}} \left( \sum_a X^A_a \partial_a
\ensuremath{\mathcal{K}} \ \phi^a + \xi_A \right)^2 \end{equation} where $X^A_a$ are Killing
vectors for $U(1)_A$. In addition, clearly the perturbative part of
the superpotential must be $U(1)_A$ invariant. But moreover, it
constrains the non-perturbative superpotential as well. In
particular, if the dilaton appears in the exponent, the product
$e^{q_\Phi S} \Phi^{\delta_{GS}/2}$ is, and must also be, $U(1)_A$
invariant.
\subsection{Target space modular invariance}
\label{sec:modular}
In this section, we wish to present the modular dependence of the
gauge kinetic function, the K\"{a}hler potential, and of the
superpotential in as general a form as possible. Most studies in
the past have worked with a universal $T$ modulus, and neglected the
effects of the $U$ moduli altogether. Such a treatment is
warranted, for example, in the $\mathbb{Z}_3$ orbifolds where there
are no $U$ moduli. If we want to work in the limit of a stringy
orbifold GUT \cite{Kobayashi:2004ya} which requires one of the $T$
moduli to be much larger than the others, or in the
\ensuremath{\mathbb{Z}_6\text{-II }} orbifolds, however, it is impossible to treat all
of the $T$ and $U$ moduli on the same footing.
Consider the $SL(2,\mathbb{Z})$ modular transformations of $T$ and
$U$ given by
\cite{Witten:1985xb,Ferrara:1986qn,Cvetic:1988yw,Shapere:1988zv,Ferrara:1989bc,Lauer:1989ax,Chun:1989se,Ferrara:1989qb,Lauer:1990tm,Erler:1991ju,Erler:1991an,Stieberger:1992bj}\footnote{For
an excellent review with many references, see \cite{Bailin:1999nk}.}
\begin{equation}
T \rightarrow \frac{aT-ib}{icT+d},\,\,\,ad-bc = 1,\,\,\,a,b,c,d\in\mathbb{Z},
\end{equation}
and
\begin{equation}
\log\left(T + \bar{T}\right) \rightarrow \log\left(\frac{T +
\bar{T}}{(icT + d)(-ic\bar{T} + d)}\right).
\end{equation}
The K\"{a}hler potential for moduli to zeroth order is given by:
\begin{eqnarray} \label{moduli_kahler0} \nonumber
\ensuremath{\mathcal{K}} &=& -\sum_{i=1}^{h_{(1,1)}} \log\left(T^i + \bar{T}^i\right) -
\sum_{j=1}^{h_{(2,1)}} \log\left(U^j + \bar{U}^j\right) \\
& = &- \sum_{i=1}^3 \log\left(T^i + \bar{T}^i\right) - \log\left(U + \bar{U}\right)
\end{eqnarray} where the last term applies to the ``mini-landscape" models, since in this case $h_{(1,1)} = 3, \
h_{(2,1)} = 1$. Under the modular group, the K\"{a}hler potential
transforms as
\begin{equation}
\ensuremath{\mathcal{K}}\rightarrow\ensuremath{\mathcal{K}} + \sum_{i=1}^{h_{(1,1)}}\log |ic_iT^i + d_i|^2 + \sum_{j=1}^{h_{(2,1)}}
\log |ic_jU^j + d_j|^2 .
\end{equation}
The scalar potential $V$ is necessarily modular invariant. We have
\begin{equation} V = e^{\cal G} \left( {\cal G}_I {\cal G}^{I \bar J} {\cal
G}_{\bar J} - 3 \right) \end{equation} where ${\cal G} = \ensuremath{\mathcal{K}} + \log|\ensuremath{\mathcal{W}}|^2$.
Hence for the scalar potential to be invariant under the modular
transformations, the superpotential must also transform as follows:
\begin{eqnarray} \nonumber \label{superpotenial_modular_transformation}
\ensuremath{\mathcal{W}} &\rightarrow& \prod_{i=1}^{h_{(1,1)}}\prod_{j=1}^{h_{(2,1)}}(ic_iT^i+d_i)^{-1}
(ic_jU^j+d_j)^{-1}\ensuremath{\mathcal{W}}, \\
\bar{\ensuremath{\mathcal{W}}} &\rightarrow& \prod_{i=1}^{h_{(1,1)}}\prod_{j=1}^{h_{(2,1)}}(-ic_i\bar{T}^i+d_i)^{-1}
(-ic_j\bar{U}^j+d_j)^{-1}\bar{\ensuremath{\mathcal{W}}}.
\end{eqnarray}
This can be guaranteed by appropriate powers of the Dedekind $\eta$
function multiplying terms in the superpotential.\footnote{These
terms arise as a consequence of world-sheet instantons in a string
calculation. In fact, world sheet instantons typically result in
more general modular functions
\cite{Lauer:1989ax,Chun:1989se,Ferrara:1989qb,Lauer:1990tm,Erler:1991ju,Erler:1991an,Stieberger:1992bj}.}
This is due to the fact that under a modular transformation, we have
\begin{equation}
\eta(T)\rightarrow(icT + d)^{1/2}\eta(T),
\end{equation}
up to a phase, where \begin{equation} \eta(T) = \exp(-\pi
T/12)\prod_{n=1}^\infty \left(1 - e^{- 2 \pi n T}\right). \label{eq:etaT}
\end{equation}
The transformation of both the matter fields and the superpotential
under the modular group fixes the modular dependence of the
interactions. A field in the superpotential transforms as
\begin{equation}
\Phi_I \rightarrow \Phi_I \prod_{i=1}^{h_{(1,1)}}\prod_{j=1}^{h_{(2,1)}}
\left(ic_iT^i+d_i\right)^{-n_I^i}\left(ic_jU^j+d_j\right)^{-\ell_I^j}.
\end{equation}
The modular weights $n_I^i$ and $\ell_I^j$
\cite{Dixon:1989fj,Ibanez:1992hc} depend on the localization of the
matter fields on the orbifold. For states $I$ in the $i$th
untwisted sector, i.e. those states with internal momentum in the
$i$th torus, we have $n_I^i = \ell_I^i = 1$, otherwise the weights
are 0. For twisted sector states, we first define $\vec{\eta}(k)$,
which is related to the twisted sector $k (=1,\ldots,N-1)$ and the
orbifold twist vector $v$ by
\begin{equation}
\eta_i(k) \equiv kv_i \ensuremath{\mathrm{~mod~}} 1.
\end{equation}
Further, we require
\begin{equation}
\sum_i \eta_i(k) \equiv 1.
\end{equation}
Then the modular weight of a state in the $k$th twisted sector is
given by
\begin{eqnarray} \label{modular_weight_1}
n_I^i \equiv & (1-\eta^i(k)) + N^i - \bar{N}^i & {\rm for} \;\; \eta_i(k) \neq 0 \\
n_I^i \equiv & N^i - \bar{N}^i & {\rm for} \;\; \eta_i(k) = 0. \nonumber
\end{eqnarray}
The $N^i \ (\bar{N}^i)$ are
\textit{integer} oscillator numbers for left-moving oscillators
$\tilde{\alpha}^i$ ($\bar{\tilde{\alpha}}^{\bar{i}}$), respectively.
Similarly,
\begin{eqnarray}
\ell_I^i \equiv & (1-\eta^i(k)) - N^i + \bar{N}^i & {\rm for} \;\; \eta_i(k) \neq 0 \\
\ell_I^i \equiv & - N^i + \bar{N}^i & {\rm for} \;\; \eta_i(k) = 0. \nonumber
\end{eqnarray}
In general, one can compute the superpotential to arbitrary order in
powers of superfields by a straightforward application of the string
selection rules \cite{Hamidi:1986vh, Casas:1991ac, Erler:1992gt,
patrick_phd}. One assumes that any term not forbidden by the string
selection rules appears with order one coefficient. In practice,
even this becomes intractable quickly, and we must cut off the
procedure at some low, finite order. More detailed calculations of
individual terms give coefficients dependent on volume moduli due to
string world sheet instantons. In general the moduli dependence can
be obtained using the constraint of target space modular invariance.
Consider a superpotential term for the ``mini-landscape" models,
with three $T$ moduli and one $U$ modulus, of the form:
\begin{equation}
\ensuremath{\mathcal{W}}_3 = w_{IJK} \Phi_I \Phi_J \Phi_K.
\end{equation}
We assume that the fields $\Phi_{I,J,K}$ transform with modular
weights $n^i_{I,J,K}$ and $\ell^3_{I,J,K}$ under $T_i, \ i=1,2,3$
and $U$, respectively. Using the (net) transformation property of
the superpotential, and the transformation property of $\eta(T)$
under the modular group, we have (for non-universal moduli):
\begin{equation}
\nonumber \label{t_depend}
w_{IJK} \sim h_{IJK} \prod_{i=1}^3 \eta(T_i)^{\gamma_{T_i}} \eta(U)^{\gamma_{U}}
\end{equation} where $ \gamma_{T_i} = -2(1-n_I^i-n_J^i-n_K^i) , \; \gamma_U =
-2(1-\ell_I^3-\ell_J^3-\ell_K^3)$.\footnote{Note, the constants
$\gamma_{T_i}, \ \gamma_U$ can quite generally have either sign,
depending upon the modular weights of the fields at the particular
vertex.} This is easily generalized for higher order interaction
terms in the superpotential. We see that the modular dependence of
the superpotential is rarely symmetric under interchange of the
$T_i$ or $U_i$. Note, when minimizing the scalar potential we
shall use the approximation $\eta(T)^{\gamma_T} \approx e^{-bT}$
with $b = \pi \gamma_T/12$. (Recall, at large $T$, we have
$\log(\eta(T)) \approx - \pi T/12$.) This approximation misses the
physics near the self-dual point in the potential, nevertheless, it
is typically a good approximation.
As a final note, Wilson lines break the $SL(2, \mathbb{Z})$ modular
group down to a subgroup \cite{Love:1996sk} (see Appendix
\ref{app:diff_racetrack}). This has the effect of an additional
differentiation of the moduli as they appear in the superpotential.
In particular, factors of $\eta(T_i)$ are replaced by factors of
$\eta(N T_i)$ or $\eta(T_i/N)$ for Wilson lines in $\mathbb{Z}_N$.
In summary, the different modular dependence of twisted sector
fields and the presence of Wilson lines leads quite generally to
anisotropic orbifolds \cite{Bailin:1994hu}.
\subsection{Gauge kinetic function and sigma model anomaly}
\label{sec:gaugekinetic}
To one loop, the string-derived gauge kinetic function is given by
\cite{Dixon:1990pc,Derendinger:1991hq,Ibanez:1991zv,Lust:1991yi,Ibanez:1992hc,Kaplunovsky:1995jw}
\begin{eqnarray} \nonumber \label{gauge_kinetic_one_loop}
f_a(S,T) &=& k_a S + \frac{1}{8\pi^2} \sum_{i=1}^{h_{(1,1)}}
\left(\alpha_a^i - k_a\delta_{\sigma}^i\right) \log \left(\eta(T^i)\right)^2 \\
&& + \frac{1}{8\pi^2} \sum_{j=1}^{h_{(2,1)}}
\left(\alpha_a^j - k_a\delta_{\sigma}^j\right) \log \left(\eta(U^j)\right)^2
\end{eqnarray}
where $k_a$ is the Ka\v{c}-Moody level of the group, which we will
normally take to be 1. The constants $\alpha_a^i$ are model
dependent, and are defined as
\begin{equation} \nonumber
\alpha_a^i \equiv \ell(\text{adj})- \sum_{\text{rep}_I} \ell_a(\text{rep}_I)(1+2n^i_I).
\end{equation}
$\ell(\text{adj})$ and $\ell_a(\text{rep}_I)$ are the Dynkin indices
of the adjoint representation and of the matter representation $I$
of the group $\ensuremath{\mathcal{G}}_a$, respectively \cite{Slansky:1981yr} and $n^i_I$
are modular weights.\footnote{If $T_a^r$ are the generators of the
group $G_a$ in the representation $r$, then we have $Tr(T_a^r T_b^r)
= \ell_a(\text{rep}_r)\delta_{a b}$.} The $\delta_{\sigma}^i$ terms
are necessary to cancel an anomaly in the underlying $\sigma$-model,
which induces a transformation in the dilaton field under the
modular group:
\begin{equation}
S \rightarrow S + \frac{1}{8\pi^2} \sum_{i=1}^{h_{(1,1)}} \delta_{\sigma}^i
\log\left(ic_iT_i + d_i\right) + \frac{1}{8\pi^2} \sum_{j=1}^{h_{(2,1)}}
\delta_{\sigma}^i \log\left(ic_jU^j+ d_j\right).
\end{equation}
It is important to note that the factor \begin{equation} \left(\alpha_a^i -
k_a\delta_{\sigma}^i\right) \equiv \frac{b_a^{(N=2)}(i)}{|D|/|D_i|}
\end{equation} where $b_a^{(N=2)}(i)$ is the beta function coefficient for the
$i$th torus. It is non-zero if and only if the $k$-th twisted
sector has an effective $N=2$ supersymmetry. Moreover this occurs only
when, in the $k$-th twisted sector, the $i$th torus is not rotated.
The factors $|D|, \ |D_i|$ are the degree of the twist group $D$ and
the little group $D_i$, which does not rotate the $i$th torus. For
example, for the ``mini-landscape" models with $D = \ensuremath{\mathbb{Z}_6\text{-II }}$
we have $|D| = 6$ and $|D_2|= 2, \ |D_3| = 3$ since the little group
keeping the second (third) torus fixed is $\mathbb{Z}_2 \
(\mathbb{Z}_3)$. The first torus is rotated in all twisted sectors.
Hence, the gauge kinetic function for the ``mini-landscape" models
is only a function of $T_2$ and $T_3$.
Taking into account the sigma model anomalies, the heterotic string
K\"{a}hler potential has the following form, where we have included
the loop corrections to the dilaton
\cite{Dixon:1990pc,Derendinger:1991hq}
\begin{eqnarray} \label{moduli_kahler} \nonumber
\ensuremath{\mathcal{K}} &=& -\log\left(S + \bar{S} + \frac{1}{8\pi^2}\sum_{i=1}^{h_{(1,1)}}
\delta_{\sigma}^i \log\left(T^i + \bar{T}^i\right) + \frac{1}{8\pi^2}
\sum_{j=1}^{h_{(2,1)}} \delta_{\sigma}^j \log\left(U^j + \bar{U}^j\right)\right) \\
&&-\sum_{i=1}^{h_{(1,1)}} \log\left(T^i + \bar{T}^i\right) -
\sum_{j=1}^{h_{(2,1)}} \log\left(U^j + \bar{U}^j\right).
\end{eqnarray}
The first line of Eqn. (\ref{moduli_kahler}) is modular invariant by
itself, and one can redefine the dilaton, $Y$, such that
\begin{equation}
Y \equiv S + \bar{S} + \frac{1}{8\pi^2}\sum_{i=1}^{h_{(1,1)}} \delta_{\sigma}^i
\log\left(T^i + \bar{T}^i\right) + \frac{1}{8\pi^2}\sum_{j=1}^{h_{(1,2)}}
\delta_{\sigma}^j \log\left(U^j + \bar{U}^j\right),
\end{equation}
where $Y$ is invariant under the modular transformations.
\subsection{Non-perturbative superpotential}
\label{sec:NP}
In all ``mini-landscape" models \cite{Lebedev:2006tr}, and most
orbifold heterotic string constructions, there exists a hidden
sector with non-Abelian gauge interactions and vector-like matter
carrying hidden sector charge. In the ``benchmark"
models \cite{Lebedev:2007hv} the hidden sector gauge group is $SU(4)$
with chiral matter in the $4 + \bar 4$ representation.
In this section let us consider a generic hidden sector with gauge
group $SU(N_{1})\otimes SU(N_2)\otimes U(1)_A$, where `$A$' stands
for anomalous. There are $N_{f_1}$ and $N_{f_2}$ flavors of quarks
$Q_{1}$ and $Q_{2}$ in the fundamental representation (along with
anti-quarks $\tilde{Q}_1$ and $\tilde{Q}_2$, in the anti-fundamental
representations), as well as two singlet fields, called $\phi$ and
$\chi$. The charge assignments are listed in Table
\ref{tab:GS_charge_assignments}. We assume the existence of two
moduli, $S$ and $T$, which enter the non-perturbative superpotential
through the gauge kinetic function, namely $f=f(S,T)$. The model
also allows for $T$ dependence in the Yukawa sector.
Non-perturbative effects generate a potential for the $S$ and $T$
moduli. Gaugino condensation will generate a scale $\ensuremath{\Lambda_{\textsc{sqcd}}}$, which
is determined purely by the symmetries of the low energy theory:
\begin{equation} \label{gaugino_condensate_general}
\Lambda_a(S,T) = e^{-\frac{8\pi^2}{\beta_a}f_a(S,T)},
\end{equation}
where $\beta_a = 3 N_a - N_{f_a}$ is the one loop beta function
coefficient of the theory. At tree level $f_a(S,T) = S$, however, we
include the possibility of threshold corrections which introduce a
dependence on the $T$ modulus
\cite{Dixon:1990pc,Derendinger:1991hq}. We also find that $U(1)_A$
and modular invariance together dictate a very specific form for the
non-perturbative superpotential.
\begin{table}
\caption{\label{tab:GS_charge_assignments} Charge assignments for the
fields in a generic hidden sector. Flavor indices are suppressed.}
\centering
\begin{tabular}{c|cccccc}
&$\phi$ &$\chi$ &$Q_1$ &$Q_2$ &$\tilde{Q}_1$ &$\tilde{Q}_2$ \\
\hline
$\U1_A$ &-1 &$q_{\chi}$ &$q_1$ &$q_2$ &$\tilde{q}_1$ &$\tilde{q}_2$ \\
$\SU{N_1}$ &1 &1 &$\Box$ &1 &$\bar{\Box}$ &1 \\
$\SU{N_2}$ &1 &1 &1 &$\Box$ &1 &$\bar{\Box}$
\end{tabular}
\end{table}
In the ``mini-landscape" analysis the effective mass terms for the
vector-like exotics were evaluated. They were given as a polynomial
in products of chiral MSSM singlet fields [chiral moduli]. It was
shown that all vector-like exotics obtain mass~\footnote{In fact,
one of the $SU(4)$ quark- anti-quark pairs remained massless in the
two ``benchmark" models.} when the chiral moduli obtain VEVs at
supersymmetric points in moduli space. In our example let
us, for simplicity, take couplings between the quarks and the field
$\phi$ to be diagonal in flavor space. Mass terms of the form \begin{equation}
\mathbb{M}_1(\phi,T) Q_{1}\tilde{Q}_{1} + \mathbb{M}_2(\phi,T)
Q_{2}\tilde{Q}_{2} \end{equation} are dynamically generated when $\phi$
receives a non-zero VEV, which we will discuss below. A key
assumption is that those mass terms are larger than the scale of
gaugino condensation, so that the quarks and anti-quarks may be
consistently integrated out. If this can be accomplished, then one
can work in the pure gauge limit
\cite{deCarlos:1991gq}.\footnote{There is a check on the consistency
of this approach: at the end of the day, after calculating the VEVs
of the scalars, we can verify that the mass terms for the
quarks are indeed of the correct magnitude.}
Before we integrate out the meson fields, the non-perturbative
superpotential (plus quark masses) for $N_{f_a} < N_a$ is of the
form \cite{Affleck:1983mk}
\begin{equation} \label{full_np_w}
\mathcal{W}_{\textsc{np}} = \sum_{a= 1,2} \left[ \mathbb{M}_a(\phi,T) Q_{a}\tilde{Q}_{a}
+ (N_a-N_{f_a})\left(\frac{\Lambda_a^{3N_a-N_{f_a}}}{\det Q_a
\tilde{Q}_a}\right)^{\frac{1}{N_a-N_{f_a}}} \right],
\end{equation}
with $\mathbb{M}_a(\phi,T) = c_a e^{-b_a T}\phi^{q_a+\tilde{q}_a}$
where $c_a$ is a constant. Note, given the charges for the fields
in Table \ref{tab:GS_charge_assignments} and using Eqns.
(\ref{eq:u1A}), (\ref{eq:GS}) and (\ref{gaugino_condensate_general}), one
sees that $\mathcal{W}_{\textsc{np}}$ is $U(1)_A$ invariant. The
K\"{a}hler potential for the hidden sector is assumed to be of the
form
\begin{eqnarray}
\ensuremath{\mathcal{K}} & = -\log(S + {\overline S}) - 3 \log(T + {\overline T}) + \alpha_\phi
{\overline \phi} e^{-2 V_A} \phi + \alpha_\chi {\overline \chi} e^{2 q_\chi V_A} \chi & \\
& + \sum_{a = 1,2} \alpha_a ({\overline Q}_a e^{V_a + 2 q_a V_A} Q_a +
{\overline {\tilde Q}}_a e^{- V_a + 2 \tilde q_a V_A} \tilde Q_a \nonumber
\end{eqnarray}
The quantities $\alpha_\phi, \alpha_\chi, \alpha_i$ are generally
functions of the modulus $T$, where the precise functional
dependence is fixed by the modular weights of the fields (see
Section \ref{sec:modular}). $V_i$ and $V_A$ denote the vector
superfields associated with the gauge groups $\ensuremath{\mathcal{G}}_i = SU(N_i)$ and
$U(1)_A$.
The determinant of the quark mass matrix is given by
\begin{equation} \label{quark_mass_matrix}
\det\mathbb{M}_a(\phi,T) = \left(c_a e^{-b_a T}\phi^{q_a+\tilde{q}_a}\right)^{N_{f_a}}.
\end{equation}
We have taken the couplings between $\phi$ and the quarks to have
exponential dependence on the $T$ modulus, an ansatz which is
justified by modular invariance (see Section \ref{sec:modular}).
Inserting the meson equations of motion and Eqn.
(\ref{quark_mass_matrix}) into Eqn. (\ref{full_np_w}), we have
\begin{equation} \nonumber \label{racetrack}
\mathcal{W}_{\textsc{np}} = \sum_{a = 1,2} \left[ N_a \left(c_a e^{-b_a T}
\phi^{q_a+\tilde{q}_a}\right)^{\frac{N_{f_a}}{N_a}}
\left[\Lambda_a(S,T)\right]^{\frac{3N_a-N_{f_a}}{N_a}} \right].
\end{equation}
Note that the transformation of the superpotential under the modular
group in Eqn. (\ref{superpotenial_modular_transformation}) also
requires that the (non-perturbative) superpotential obey
\begin{equation}
\ensuremath{\mathcal{W}}_{\textsc{np}} \rightarrow \prod_{i=1}^{h_{(1,1)}}\prod_{j=1}^{h_{(2,1)}}
(ic_iT^i+d_i)^{-1}(ic_jU^j+d_j)^{-1}\ensuremath{\mathcal{W}}_{\textsc{np}}.
\end{equation}
Because the non-perturbative lagrangian must be invariant under all
of the symmetries of the underlying string theory, it must be that
\cite{Font:1990nt,Ferrara:1990ei,Nilles:1990jv,Casas:1990qi,Lust:1991yi,deCarlos:1992da}:
\begin{equation} \label{duality_invariant_gaugino_condensate}
\ensuremath{\mathcal{W}}_{\textsc{np}} \equiv A \times e^{-a S}\prod_{i=1}^{h_{(1,1)}}
\prod_{j=1}^{h_{(2,1)}} \left(\eta(T^i)\right)^{-2+\frac{3}{4\pi^2 \beta}\delta^i_{\sigma}}
\left(\eta(U^j)\right)^{-2+\frac{3}{4\pi^2 \beta}\delta^j_{\sigma}}
\end{equation} where $a \equiv \frac{24 \pi^2}{\beta}$ and
$\beta = 3 \ell(\text{adj}) - \sum_{I} \ell(\text{rep}_I)$ is the
one-loop beta function coefficient, and $A$ is generally a function of
the chiral matter fields appearing in $\mathbb{M}$. This, coupled with the one loop
gauge kinetic function in Eqn. (\ref{gauge_kinetic_one_loop}), gives
the heterotic generalization of the Racetrack superpotential.
In the following Section \ref{sec:one_gaugino_condensate}, we
construct a simple model using the qualitative features outlined in
this section. This model is novel because it requires only one
non-Abelian gauge group to stabilize moduli and give a de Sitter
vacuum. We have also constructed two condensate models,
however, the literature already contains several examples of the
``racetrack" in regards to stabilization of $S$ and $T$ moduli.
Moreover in the ``mini-landscape" models, whose features we are
seeking to reproduce, there are many examples of hidden sectors
containing a single non-Abelian gauge group \cite{Lebedev:2006tr},
while there are no examples with multiple hidden sectors.
\section{Moduli stabilization and supersymmetry\\ breaking in the bulk}
\label{sec:one_gaugino_condensate}
In this section we construct a simple, generic heterotic orbifold
model which captures many of the features discussed in Section
\ref{sec:general_structure}. In particular, it is a single gaugino
condensate model with the following fields - dilaton ($S$), modulus
($T$) and MSSM singlets ($\phi_1, \phi_2, \chi$). The model has
one anomalous $\U1_A$ with the singlet charges given by ($q_{\phi_1}
= -2, q_{\phi_2} = -9, q_\chi = 20$). The K\"{a}hler and
superpotential are given by~\footnote{The coefficient $A$ (Eqn.
(\ref{eqn:A})) is an implicit function of all other non-vanishing
chiral singlet VEVs which would be necessary to satisfy the modular
invariance constraints, i.e. $A = A(\langle \phi_I \rangle)$. If
one re-scales the $U(1)_A$ charges, $q_{\phi_i}, q_\chi \rightarrow
q_{\phi_i}/r, q_\chi/r$, then the $U(1)_A$ constraint is satisfied
with $r = 15 p$ (assuming no additional singlets in $A$). Otherwise
we may let $r$ and $p$ be independent. This re-scaling does not
affect our analysis, since the vacuum value of the $\phi_i, \chi$
term in the superpotential vanishes. \label{footnote} }
\begin{eqnarray} \label{model_Kahler_potential}
\ensuremath{\mathcal{K}} = & -\log[S + \bar S] - 3 \log[T + {\overline T}] + {\overline \phi}_1 \phi_1 +
{\overline \phi}_2 \phi_2 + {\overline \chi} \chi & \\ \label{model_superpotential}
\ensuremath{\mathcal{W}} = & e^{-b T} (w_0 + \chi (\phi_1^{10} + \lambda \phi_1 \phi_2^2))
+ A \ \phi_2^p \ e^{-a S - b_2 T}. & \label{eqn:A}
\end{eqnarray}
In addition, there is an anomalous $U(1)_A$ $D$-term given by
\begin{eqnarray}
D_A = & 20 {\overline \chi} \chi - 2 {\overline \phi}_1 \phi_1 - 9 {\overline \phi}_2 \phi_2 -
\frac{1}{2} \delta_{GS} \ \partial_S \ensuremath{\mathcal{K}} &
\end{eqnarray}
with $\delta_{GS} = \frac{(q + \tilde q) N_{f}}{4 \pi^2} = N_f/(4
\pi^2)$.
In the absence of the non-perturbative term (with coefficient $A$)
the theory has a supersymmetric minimum with $\langle \chi \rangle =
\langle \phi_1 \rangle = 0$ and $\langle \phi_2 \rangle \neq 0$ and
arbitrary. This property mirrors the situation in the
``mini-landscape" models where supersymmetric vacua have been found
in the limit that all non-perturbative effects are neglected. We
have also added a constant $w_0 = w_0(\langle \phi_I \rangle)$ which
is expected to be generated (in the ``mini-landscape" models) at
high order in the product of chiral moduli due to the explicit
breaking of an accidental $R$ symmetry which exists at lower
orders~\cite{Kappl:2008ie}.\footnote{The fields entering $w_0$ have
string scale mass.} The $T$ dependence in the superpotential is
designed to take into account, in a qualitative way, the modular
invariance constraints of Section \ref{sec:modular}. We have
included only one $T$ modulus, assuming that the others can be
stabilized near the self-dual point
\cite{Font:1990nt,Cvetic:1991qm}. Moreover, as argued earlier, the
$T_i$ and $U$ moduli enter the superpotential in different ways (see
Section \ref{sec:modular}). This leads to modular invariant
solutions which are typically
anisotropic~\cite{Bailin:1994hu}.\footnote{Note, we have chosen to
keep the form of the K\"{a}hler potential for this single $T$
modulus with the factor of 3, so as to maintain the approximate
no-scale behavior.}
Note, that the structure, $\ensuremath{\mathcal{W}} \sim w_0 e^{- b T} + \phi_2 \ e^{- a
S - b_2 T}$ gives us the crucial progress\footnote{Note, the
constants $b, \ b_2$ can have either sign. For the case with $b, \
b_2 > 0$ the superpotential for $T$ is racetrack-like. However for
$b, \ b_2 < 0$ the scalar potential for $T$ diverges as $T$ goes to
zero or infinity and compactification is guaranteed
\cite{Font:1990nt,Cvetic:1991qm}.} -
\begin{enumerate}[i.)]
\item a `hybrid KKLT' kind of superpotential that behaves like a
single-condensate for the dilaton $S$, but as a racetrack for the
$T$ and, by extension, also for the $U$ moduli; and
\item an additional matter $F_{\phi_2}$ term driven by the
cancelation of the anomalous $U(1)_A$ $D$-term seeds SUSY breaking
with successful uplifing.
\end{enumerate}
The constant $b$ is fixed by modular invariance constraints. In
general the two terms in the perturbative superpotential would have
different $T$ dependence. We have found solutions for this case as
well. This is possible since the VEV of the $\chi$ term in the
superpotential vanishes. The second term (proportional to $A$)
represents the non-perturbative contribution of one gaugino
condensate. The constants $a = 24 \pi^2/\beta, \ b_2$ and $p$
depend on the size of the gauge group, the number of flavors and the
coefficient of the one-loop beta function for the effective $N=2$
supersymmetry of the torus $T$. For the ``mini-landscape" models,
this would be either $T_2$ or $T_3$. Finally, the coefficient of the
exponential factor of the dilaton $S$ is taken to be $A \ \phi_2^p$.
This represents the effective hidden sector quark mass term, which
in this case is proportional to a power of the chiral singlet
$\phi_2$. In a more general case, it would be a polynomial in powers
of chiral moduli.\footnote{Holomorphic gauge invariant monomials
span the moduli space of supersymmetric vacua. One such monomial is
necessary to cancel the Fayet-Illiopoulos $D$-term (see Appendix
\ref{app:HIM}).} The exponent $p$ depends in general on the size of
the gauge group, the number of flavors and the power that the field
$\phi_2$ appears in the effective quark mass term.
We have performed a numerical evaluation of the scalar potential
with the following input parameters. We take hidden sector gauge
group $SU(N)$ with $N = 5, \ N_f = 3$ and $a = 8
\pi^2/N$.\footnote{We have also found solutions for the case with $N
= 4, \ N_f = 7$ which is closer to the ``mini-landscape" benchmark
models. Note, when $N_f > N$ we may still use the same formalism,
since we assume that all the $Q, \tilde Q$s get mass much above the
effective QCD scale.} For the other input values we have considered
five different possibilities given in Table
\ref{tab:onecond.input}.\footnote{Note the parameter relation $r =
15 p$ in Table \ref{tab:onecond.input} is derived using $U(1)_A$
invariance and the assumption that no other fields with
non-vanishing $U(1)_A$ charge enter into the effective mass matrix
for hidden sector quarks. We have also allowed for two cases where
this relation is not satisfied.} We find that supersymmetry
breaking, moduli stabilization and up-lifting is a direct
consequence of adding the non-perturbative superpotential term.
In our analysis we use the scalar potential $V$ given by \begin{equation} V =
e^K (\sum_{i = 1}^5\ \sum_{j = 1}^5 \left[ F_{\Phi_i}\
\overline{F_{\Phi_j}}\ \ensuremath{\mathcal{K}}^{-1}_{i, j} - 3 |W|^2 \right]) +
\frac{D_A^2}{(S + \bar S)} + \Delta V_{CW}[\Phi_i,
\overline{\Phi}_i] \end{equation} where $\Phi_{i,j} = \{ S, T, \chi, \phi_1,
\phi_2 \}$ and $F_{\Phi_i} \equiv \partial_{\Phi_i} \ensuremath{\mathcal{W}} +
(\partial_{\Phi_i} \ensuremath{\mathcal{K}}) \ensuremath{\mathcal{W}}$. The first two terms are the tree
level supergravity potential. The last term is a one loop
correction which affects the vacuum energy and $D$ term
contribution.
The one loop Coleman-Weinberg potential is in general given by \begin{equation}
\Delta V_{CW} = \frac{1}{32 \pi^2} Str(M^2) \Lambda^2 + \frac{1}{64
\pi^2} Str(M^4 \log[\frac{M^2}{\Lambda^2}]) \end{equation} with the mass
matrix $M$ given by $M = M(\Phi_i)$ and $\Lambda$ is the relevant
cut-off in the problem. We take $\Lambda = M_{\textsc{s}} \sim 10^{17}$
GeV.
We have not evaluated the full one loop correction. Instead we use
the approximate formula \begin{equation} \Delta V_{CW}[\phi_2, \overline{\phi_2}]
= \frac{\lambda^2 \ F_2^2 \ |\phi_2|^2}{8 \pi^2} \left(\log[R
(\lambda |\phi_2|^2)^2] + 3/2\right) + \order{\Lambda^2} \end{equation} where
$F_2 = \langle F_{\phi_2} \rangle$ is obtained self-consistently and
all dimensionful quantities are expressed in Planck units. This one
loop expression results from the $\chi, \ \phi_1$ contributions to
the Coleman-Weinberg formula. The term quadratic in the cut-off is
naturally proportional to the number of chiral multiplets in the
theory and could be expected to contribute a small amount to the
vacuum energy, of order a few percent times $m_{3/2}^2 M_{pl}^2$. We
will discuss this contribution later, after finding the minima of
the potential. Finally, note that the parameters $\lambda, \ R$ in
Table \ref{tab:onecond.input} might both be expected to be
significantly greater than one when written in Planck units. This
is because the scale of the effective higher dimensional operator
with coefficient $\lambda$ in Eqn. \ref{eqn:A} is most likely set by
some value between $M_{Pl}$ and $M_{string}$ and the cut-off scale
for the one loop calculation (which determines the constant $R$) is
the string scale and not $M_{Pl}$.
\begin{table}[ht]
\renewcommand\arraystretch{1.2}
\begin{center}
\begin{tabular}{|c|c|c|c|c|c|c|c|c|}
\hline
Case & $b$ & $b_2$ & $\lambda$ & $R$ & $p $ & $r$ & $A$ & $w_0$ \\
\hline
1 & $\pi/50$ & $3\pi/2$ & $33$ & $10$ & $2/5$ & $15p$ & $160$ & $8 \times 10^{-15}$ \\
\hline
2 & $8/125$ & $3\pi/2$ & $0$ & $5$ & $2/5$ & $15p$ & $30$ & $42 \times 10^{-16}$ \\
\hline
3 & $1/16$ & $29\pi/20$ & $38$ & $10$ & $2/5$ & $15p$ & $90$ & $6 \times 10^{-15}$ \\
\hline
4 & $-\pi/120$& $-\pi/40$ & $40$ & $64$ & $2/3$ & $1$ & $1/10$& $-5 \times 10^{-15}$ \\
\hline
5 & $-\pi/250$ & $-\pi/100$ & $25$ & $16$ & $1$ & $10/3$& $7/5$ & $-7 \times 10^{-15}$ \\
\hline
\end{tabular}
\caption{\label{tab:onecond.input}Input values for the
superpotential parameters for three different cases. Case 2 has a
vanishing one loop correction for $\phi_2$.}
\end{center}
\end{table}
\begin{table}[ht]
\renewcommand\arraystretch{1.2}
\centering
\noindent\makebox[\textwidth]{ \scalebox{0.75}{
\begin{tabular}{|c|c|c|c|c|c|}
\hline
& Case 1 & Case 2 & Case 3 & Case 4 & Case 5\\
\hline
$\vev{s}$ & $2.2$ & $2.2$ & $2.1$ & $2.1$ & $2.2$ \\
\hline
$\vev{t}$ & $1.2$ & $1.1$ & $1.6$ & $1.1$ & $1.1$ \\
\hline
$\vev{\sigma}$ & $1.0$ & $1.0$ & $1.0$ & $0.0$ & $0.0$ \\
\hline
$\vev{\phi_2}$ & $0.08$ & $0.08$ & $0.08$ & $0.03$ & $0.06$ \\
\hline
$F_S$ & $2.8 \times 10^{-16}$ & $1.3 \times 10^{-16}$ & $2.7 \times 10^{-16}$ & $1.1 \times 10^{-16}$ & $8.0\times10^{-17}$\\
\hline
$F_T$ & $-8.7 \times 10^{-15}$ & $-5.1 \times 10^{-15}$ & $-5.0 \times 10^{-15}$ & $6.7 \times 10^{-15}$ & $9.1\times10^{-15}$\\
\hline
$F_{\phi_2}$ & $-9.2\times 10^{-17}$ & $-4.5\times 10^{-17}$ & $-8.9\times 10^{-17}$ & $1.3\times10^{-15}$ & $1.3\times10^{-15}$ \\
\hline
$D_A $ & $4.4\times 10^{-31}$ & $1.0\times 10^{-32}$ & $5.9\times 10^{-31}$ & $-3.8\times 10^{-31}$ & $-4.8\times10^{-32}$ \\
\hline
$D_A / m_{3/2}^2$ & $0.6$ & $0.03$ & $2.7$ & $-0.7$ & $-0.05$ \\
\hline
$V_0/(3 m_{3/2}^2)$ & $-0.02 $ & $-0.01 $ & $-0.02$ & $-0.03$ & $-0.02$\\
\hline
$m_{3/2}$ & $2.2$ TeV & $1.4$ TeV & $1.1$ TeV & $1.8$ TeV & $2.4$ TeV \\
\hline
\end{tabular}
}}
\caption{\label{tab:onecond.result} The values for field VEVs and
soft SUSY breaking parameters at the minimum of the scalar potential.
Note $F_\Phi \equiv \partial_\Phi \ensuremath{\mathcal{W}} + (\partial_\Phi \ensuremath{\mathcal{K}}) \ensuremath{\mathcal{W}}$.}
\end{table}
In all cases we find a meta-stable minimum with all (except for two
massless modes) fields massive of \order{\text{TeV}} or larger.
Supersymmetry is broken at the minimum with values given in Table
\ref{tab:onecond.result}. Note $\re{S} \sim 2.2$ and $\re{T}$ ranges
between 1.1 and 1.6. The moduli $\chi, \ \phi_1$ are stabilized at
their global minima $\phi_1 = \chi = 0 $ with $F_\chi = F_{\phi_1} =
0$ in all cases. The modulus $\sigma = \im{S}$ is stabilized at
$\sigma \approx 1$ in the racetrack cases 1,2,and 3. This value
enforces a relative negative sign between the two terms dependent on
$\re{T}$. We plot the scalar potential $V$ in the \re{T} direction
for case 2 ($b, \ b_2 > 0$) (Fig. \ref{fig:pos_b}) and for case 4
($b, \ b_2 < 0$) (Fig. \ref{fig:neg_b}). Note the potential as a
function of \re{S} is qualitatively the same for both cases (Fig.
\ref{fig:v_of_s}).
\begin{figure}[t!]
\centering
\subfigure[The scalar potential in Case 2 for ${\rm Re}\,T$, with $b_i > 0$.]
{\includegraphics[scale=0.8]{pos_b.eps} \label{fig:pos_b} }
\subfigure[The scalar potential in Case 4 for ${\rm Re}\,T$, with $b_i < 0$.]
{\includegraphics[scale=0.8]{neg_b.eps} \label{fig:neg_b} }
\label{fig:potentials}
\caption[]{As ${\rm Re}\,T \rightarrow \infty$, the potential for
$b_i > 0$ mimics a Racetrack, which can be seen from Eqn.
(\ref{model_superpotential}), for example. In the case where
$b_i < 0$, however, the potential exhibits a different asymptotic
behavior. As ${\rm Re}\,T \rightarrow \infty$ the potential
diverges, which means that theory is forced to be compactified
\cite{Font:1990nt,Cvetic:1991qm}. }
\end{figure}
\begin{figure}[t!]
\centering
{\includegraphics[scale=0.8]{v_of_s} }
\caption[]{The scalar potential in the \re{S} direction for Case 2. \label{fig:v_of_s}}
\end{figure}
\begin{figure}[t!]
\centering
\includegraphics[scale=0.8]{one_loop}
\label{fig:one_loop_potential}
\caption{The one loop Coleman-Weinberg potential (Case 4) for $\phi_2$.
The dashed line represents the VEV of $\phi_2$ in the minimum of the
\textit{full} potential.}
\end{figure}
At the meta-stable minimum of the scalar potential we find a vacuum
energy which is slightly negative, i.e. of order $(-0.03 \; {\rm to}
\; -0.01)\times 3 m_{3/2}^2 M_{Pl}^2$ (see Table
\ref{tab:onecond.result}). Note, however, one loop radiative
corrections to the vacuum energy are of order $(N_T \ m_{3/2}^2
M_S^2/16 \pi^2)$ ,where $N_T$ is the total number of chiral
multiplets~\cite{Ferrara:1994kg} and we have assumed a cut-off at
the string scale $M_S$. With typical values $N_T \sim \order{300}$
and $M_S/M_{Pl} \sim 0.1$, this can easily lift the vacuum energy
the rest of the way to give a small positive effective cosmological
constant which is thus a meta-stable local dS minimum. Note that
the constants $\lambda, \ R$ have also been used to adjust the value
of the cosmological constant as well as, and more importantly for
LHC phenomenology, the value of $D_A$ (see Fig.
\ref{fig:one_loop_potential}).
The two massless fields can be seen as the result of two $U(1)$
symmetries; the first is a $U(1)_R$ symmetry and the second is
associated with the anomalous $U(1)_A$. The $U(1)_R$ is likely
generic (but approximate), since even the ``constant" superpotential
term needed to obtain a small cosmological constant necessarily
comes with $\eta(T)$ moduli dependence. Since we have approximated
$\eta(T) \sim \exp(- \pi T/12)$ by the first term in the series
expansion (Eqn. \ref{eq:etaT}), the symmetry is exact. However
higher order terms in the expansion necessarily break the $U(1)_R$
symmetry. The $U(1)_A$ symmetry is gauged.
One can express the fields $S,T,$ and $\phi_2$ in the following
basis\footnote{The fields $\chi$ and $\phi_1$ cannot be expressed in
polar coordinates as they receive zero VEV, and cannot be
canonically normalized in this basis.}:
\begin{eqnarray} \nonumber
S &\equiv& s + i\sigma, \\
T &\equiv& t+i\tau, \\ \nonumber
\phi_2 &\equiv& \varphi_2 \ e^{i\theta_2}.
\end{eqnarray}
The transformation properties of the fields $\sigma, \tau$ and
$\theta_2$ under the two \U1's are given by
\begin{eqnarray} \nonumber
U(1)_R &:& \left\{ \begin{array}{rcl} \tau & \rightarrow & \tau + c \\ \sigma &\rightarrow& \sigma + \frac{-b_2 + b}{a} c \end{array} \right. , \\
U(1)_A &:& \left\{ \begin{array}{rcl} \theta & \rightarrow & \theta -\frac{9}{r}c' \\ \sigma &\rightarrow& \sigma - \frac{9p}{a\cdot r} c' \end{array} \right. ,
\end{eqnarray}
where $c,c'$ are arbitrary constants and for the definition of $r$
see footnote \ref{footnote}. The corresponding Nambu-Goldstone (NG)
bosons are given by
\begin{eqnarray} \nonumber
\chi^1_{\textsc{ng}} &=& \frac{a}{-b_2 + b} \sigma + \tau, \\
\chi^2_{\textsc{ng}} &=& \tilde{N}\left(-\sigma + \frac{-b_2 + b}{a}\tau\right) + \frac{1}{p}\theta_2,
\end{eqnarray}
where $\tilde{N}$ is a normalization factor. One can then calculate
the mass matrix in the $\sigma-\tau-\theta_2$ basis and find two
zero eigenvalues (as expected) and one non-zero eigenvalue. The two
NG modes, in all cases, can be shown to be linear combinations of
the two eigenvectors of the two massless states. The $\U1_A$ NG
boson is eaten by the $\U1_A$ gauge boson, while the $\U1_R$
pseudo-NG boson remains as an ``invisible axion"
\cite{Nilles:1981py}. The $U(1)_R$ symmetry is non-perturbatively
broken (by world-sheet instantons) at a scale of order \begin{equation} \langle
e^{\ensuremath{\mathcal{K}}/2} \ensuremath{\mathcal{W}} \ e^{- \pi T} \rangle \approx m_{3/2} \langle e^{-
\pi T} \rangle \sim 0.02 \ m_{3/2} \end{equation} in Planck units, resulting in
an ``axion" mass of order 10 GeV and decay constant of order
$M_{Pl}$.\footnote{In addition, the heterotic orbifold models might
very well have the standard invisible axion \cite{Choi:2009jt}.}
Before discussing the rest of the moduli, in a more complete string
model, and how they would be stabilized or the LHC phenomenology of
the mini-version of the mini-landscape models, it is worth comparing
our analysis with some previous discussions in the literature.
In a series of two papers by Dvali and Pomarol
\cite{Dvali:1996rj,Dvali:1997sr}, the authors consider an anomalous
\U1 with two charged singlet fields. The $D$ term is given
by\footnote{We refer to the anomalous $U(1)$ as $U(1)_A$ and not
$U(1)_X$, as in the papers referenced below.}
\begin{equation}
D_A = q |\phi_+|^2 - |q_-|^2 + \xi
\end{equation}
The gauge invariant superpotential is
\begin{equation}
\ensuremath{\mathcal{W}} = m \phi_+ \phi_-,
\end{equation}
where $m$ has some charge under $\U1_A$. They suggest a few
different ways to generate $m$. The first is with some high power of
one of the $\phi$ fields:
\begin{equation}
\ensuremath{\mathcal{W}} \sim \phi_-^q \phi_+ \Rightarrow m \equiv \vev{\phi_-}^{q-1}
\end{equation}
The second is by giving the $\phi$ a coupling to some quarks from a
SUSY QCD theory that becomes strongly coupled. The scale,
$\Lambda_{SQCD}$ then serves as the mass term in the superpotential.
They do not, however, consider dilaton dependence, and their $D$
term is static, not dynamic. They also work in the global SUSY
limit, so they do not consider up-lifting.
In a paper by Binetruy and Dudas \cite{Binetruy:1996uv}, the authors
assume that $S$ can be stabilized at some finite value $S_0$,
possibly through some extra $S$ dependent term in the superpotential
and they assume that $F_S(S_0) = 0$. In their setup, they have an
anomalous $\U1$, some charged singlets, and some hidden sector SQCD
with matter. The singlets couple to matter, and SQCD becomes
strongly coupled, generating a scale, just as in our analysis. Since
they are working in the global SUSY limit, they are not concerned
with up-lifting.
Lalak \cite{Lalak:1997ar} considers several types of models with an
anomalous \U1, some charged singlets, and some coupling to the
dilaton $S$. In the last section, he considers superpotentials with
an exponential dependence on $S$. He then assumes that $S_0$ is a
(globally) supersymmetric minimum of the potential. Also, working in
global SUSY, he does not address up-lifting.
In a paper by Dudas and Mambrini \cite{Dudas:2006vc}, the authors
consider one modulus, one singlet field, and an $SU(N)$ with one
flavor of quarks. The $SU(N)$ becomes strongly coupled, and the
superpotential and K\"{a}hler potential look like: \begin{eqnarray} \ensuremath{\mathcal{W}} = w_0 +
(c / X^{2}) e^{-a T} + m \phi^q X \\ \ensuremath{\mathcal{K}} = -3 \log( T + \bar T -
|X|^2 - |\phi|^2 ) \end{eqnarray} where $X$ is the meson field and $\phi$ is
the singlet. Note, the modulus appearing in the exponent is $T$,
not $S$. They find that the only consistent minimum with
approximately zero cosmological constant requires $m_{3/2} \sim
\xi$. So either the gravitino mass is of order the GUT scale or for
the gravitino mass of order a TeV, the meson charge must satisfy $q
\sim 10^{-8}$.
In a paper by Dudas et al. \cite{Dudas:2007nz}, the authors consider
a single modulus and two singlet fields:
\begin{eqnarray}
D_A &=& |\phi_+|^2 - |\phi_-|^2 + \xi, \\
\ensuremath{\mathcal{W}} &=& w_0 + m\phi_+\phi_- + a \phi_-^q e^{-bT}.
\end{eqnarray}
They do not discuss the origin of the constant $w_0$. They suggest
that $m$ might come from non-perturbative effects. Note the latter
is crucial, since $m$ affects the up-lifting of the scalar
potential. They are also interested in large volume
compactifications, as $t\equiv \re{T}\approx 60$. Given their SUSY
breaking scheme, they go on to look at the low energy spectrum.
However, they neglect the $D$ term contributions to the soft masses,
claiming that there are only two possibilities for the low energy
physics:
\begin{itemize}
\item Because $\xi > 0$, some SM quarks and leptons carry
positive $\U1_A$ charges. This leads to scalar masses (for
them) of around 100 TeV, and may give an unstable low energy
spectrum.
\item All SM quarks and leptons are neutral under $\U1_X$.
This implies that there should be more matter that is charged
under the MSSM and $\U1_A$.
\end{itemize}
It seems that they have missed an important possibility, namely that
matter in the MSSM appears with $\U1_A$ charges of both signs. This
actually seems to be the generic case, at least in the
mini-landscape models.
The last paper we consider, by Gallego and Serone
\cite{Gallego:2008sv}, contains an analysis which is possibly most
similar to that in this paper. There are however two major
differences. If one neglects all non-perturbative dependence on the
dilaton and K\"{a}hler moduli, then their superpotential is of the
form $\ensuremath{\mathcal{W}} \supset \phi^q \chi$ and the D term is given by $D_A = q
|\chi|^2 - |\phi|^2 + \xi$. Hence the model does not have a
supersymmetric minimum in the global limit, due to a conflict
between $F_\chi = 0$ and $D_A = 0$. However in oure model (Eqn.
\ref{eqn:A}) there is a supersymmetric solution when
non-perturbative effects are ignored. Finally, the authors were not
able to find a supersymmetry breaking solution, like ours, with just
one hidden non-Abelian gauge sector.
As an aside, we note that Casas et al. \cite{Casas:1990qi} study a
similar problem of moduli stabilization and SUSY breaking, but
without the anomalous \U1. However, their model is very different
from ours, but they do include the one loop Coleman-Weinberg
corrections.
\section{Moduli stabilization continued - the twisted sector and blow-up moduli}
\label{sec:moduli_stabilization}
In our discussion above we considered a simple model which is
representative of heterotic orbifold models. Our simple model had
only a few moduli, i.e. the dilaton, $S$, a volume modulus, $T$,
and three chiral singlet `moduli', $\chi, \ \phi_1, \ \phi_2$. Any
heterotic orbifold construction, on the other hand, will have
several volume and complex structure moduli and, of order 50 to 100
chiral singlet moduli. The superpotential for the chiral singlet
moduli is obtained as a polynomial product of holomorphic gauge
invariant monomials which typically contain hundreds of terms at
each order (with the number of terms increasing with the order). In
the ``mini-landscape" analysis, supersymmetric vacua satisfying $F =
D = 0$ constraints to sixth order in chiral singlet moduli could be found.
Although there are many flat
directions in moduli space, the anomalous $D$-term fixes at least one
holomorphic gauge invariant monomial to have a large value. Our
simple model expressed this fact with the chiral singlets $\chi, \
\phi_1, \ \phi_2$, where the VEVs were fixed by the global SUSY
minimum with $\langle \phi_2 \rangle$ fixed by the $U(1)_A$ $D$-term.
In addition to the non-Abelian hidden gauge sector considered in the
simple model, a generic orbifold vacuum also has additional $U(1)$
gauge interactions and vector-like exotics which obtain mass
proportional to chiral singlet VEVs. Some of these singlets are
assumed to get large VEVs (of order the string scale). These are the
ones giving mass to the extra $U(1)$ gauge sector and vector-like
exotics. These same VEVs generate non-trivial Yukawa couplings for
quarks and leptons. Moreover, there are chiral singlets which get zero
VEVs, such as $\chi$ and $\phi_1$. For example, in the
``mini-landscape" benchmark model 1, the electroweak Higgs $\mu$
term is zero in the supersymmetric limit. The question arises as to
what happens to all these VEVs once supersymmetry is broken.
We now sketch the fact that the supersymmetry breaking discussed
above, ensuing from $F$-terms, $F_S, F_T, F_{\phi_2} \neq 0$ and
driven by the non-perturbative superpotential, inevitably leads to a
stabilization of the many singlet `moduli' of the heterotic orbifold
vacuum. We shall consider here 3 classes of heterotic MSSM singlets.
\subsection{Singlets with polynomial Yukawa couplings}
\label{subsubsec:NONflats}
Let us first consider singlets having polynomial Yukawa couplings in
the superpotential, which in case of a coupling arising among purely
untwisted sector fields $\phi_i^{(U)}$ are perturbatively generated,
and in the other case involving {\it at least one} twisted sector
field $\phi_i^{(T)}$ are non-perturbatively generated (see Section
\ref{sec:modular}). The latter case is actually the most common
situation. Restricting again for reasons of simplicity to the case
of a single scalar field of the type under consideration, we can
describe the two cases as follows:
\begin{itemize}
\item i) \[\ensuremath{\mathcal{K}}=-3\log\left(T+\bar T-\bar\phi^{(U)}\phi^{(U)}\right)\quad,\quad \ensuremath{\mathcal{W}}\supset \lambda\cdot\left(\phi^{(U)}\right)^N\;,\;N\geq3\]\\
Note that the untwisted sector scalar fields $\phi^{(U)}$,
being inherited from the bulk $\bf 248$ in 10d, appear this
way in the K\"{a}hler potential.
\item ii) \[\ensuremath{\mathcal{K}}=-3\log(T+\bar T)+c\bar\phi^{(T)}\phi^{(T)}\quad,\quad \ensuremath{\mathcal{W}}\supset e^{-b T}\left(\phi^{(T)}\right)^N\;,\;N\geq3\]\\
Here the exponential dependence on $T$ arises from the
$\eta$-function, which a non-perturbatively
generated Yukawa coupling must have for reasons of modular
invariance (see Section \ref{sec:modular}).
\item iii) \[\ensuremath{\mathcal{K}}=-3\log\left(T+\bar T-\bar\phi^{(U)}\phi^{(U)}\right)+c\,\bar\phi^{(T)}\phi^{(T)}\]\\[-4ex] \[\ensuremath{\mathcal{W}}\supset \lambda e^{- b T}\left(\phi^{(T)}\right)^{N}+\tilde\lambda e^{- \tilde b T}\left(\phi^{(T)}\right)^{\tilde N}\left(\phi^{(U)}\right)^{M}\;{\rm with}\;\;M,N,\tilde N\geq2\]\\
Here, too, the exponential dependence on $T$ from the
$\eta$-function dependence of a non-perturbatively
generated Yukawa coupling.
\end{itemize}
The calculation in case i) simplifies by the fact that there $K$
fulfills an extended no-scale relation
\begin{eqnarray} \ensuremath{\mathcal{K}}_{i}\ensuremath{\mathcal{K}}^{i\bar j}\ensuremath{\mathcal{K}}_{\bar j}&=&3\quad\forall\;i,j=T,\phi^{(U)}\nonumber\\
\ensuremath{\mathcal{K}}^i=\ensuremath{\mathcal{K}}^{i\bar j}\ensuremath{\mathcal{K}}_{\bar j}&=&-\mathcal{V}\cdot\delta^i_T\quad,\quad
\mathcal{V}\equiv\left(T+\bar T-\bar\phi^{(U)}\phi^{(U)}\right) \end{eqnarray}
which implies for the F-term scalar potential a result \begin{eqnarray}
V_F&=&e^\ensuremath{\mathcal{K}}\left[\ensuremath{\mathcal{K}}^{\phi^{(U)}\bar\phi^{(U)}}\left(|\partial_{\phi^{(U)}}\ensuremath{\mathcal{W}}|^2+\left(\partial_{\phi^{(U)}}\ensuremath{\mathcal{W}}\cdot\overline{\ensuremath{\mathcal{K}}_{\phi^{(U)}}\ensuremath{\mathcal{W}}}+c.c.\right)\right)\right.\nonumber\\
&&\qquad\left.+\frac{\mathcal{V}}{3}(T+\bar T)|\partial_T
\ensuremath{\mathcal{W}}|^2+\left(\mathcal{V}\partial_T\ensuremath{\mathcal{W}}+c.c.\right)\right]\quad. \end{eqnarray}
It is clear then that one solution to $\partial_{\phi^{(U)}}V_F=0$
is given by \begin{equation}
\partial_{\phi^{(U)}}\ensuremath{\mathcal{W}}=\partial_{\phi^{(U)}}\mathcal{V}=0\quad\Rightarrow\quad\langle\phi_{(U)}\rangle=0
\end{equation} because $\partial_{\phi^{(U)}}\partial_{T}\ensuremath{\mathcal{W}}\equiv 0\;\forall
\phi^{(U)}$. This implies that those untwisted sector singlets that
were stabilized at the origin in global supersymmetry by a purely
untwisted sector Yukawa coupling remain so even in supergravity.
For the twisted sector case ii) we find the scalar potential to be
\begin{eqnarray}
V_F&=&e^\ensuremath{\mathcal{K}}\left[\ensuremath{\mathcal{K}}^{\phi^{(T)}\bar\phi^{(T)}}|D_{\phi^{(T)}}\ensuremath{\mathcal{W}}|^2+\ensuremath{\mathcal{K}}^{T\bar
T}\left(|\partial_T
\ensuremath{\mathcal{W}}|^2+\underbrace{\partial_T\ensuremath{\mathcal{W}}}_{\sim F_T}\overline{\ensuremath{\mathcal{K}}_T\ensuremath{\mathcal{W}}}+c.c.\right)\right]\nonumber\\
&\sim&e^{- 2 b T}\left(\bar\phi^{(T)}\phi^{(T)}\right)^{N-1}-
F_T(T+\bar T)e^{-b T}\left(\phi^{(T)}\right)^N+c.c. \end{eqnarray} which gives
two solutions to $\partial_{\phi^{(T)}}V_F=0$ as \begin{equation}
\langle\phi^{(T)}\rangle=0\quad\bigvee\quad\langle\phi^{(T)}\rangle\sim\left(\frac{F_T
(T+\bar T)}{e^{-b
T}}\right)^{\frac{1}{N-2}}\sim\left(\frac{m_{3/2}}{e^{-b
T}}\right)^{\frac{1}{N-2}}\quad. \end{equation} This implies that the
$\phi^{(T)}$ get stabilized either at the origin, or at non-zero but
small VEVs $\ll 1$. Their value in the latter case approaches
$\phi^{(T)}\sim M_{\rm GUT}$ for non-perturbative Yukawa couplings
of order $N\gtrsim 5$ and $m_{3/2}\sim {\rm TeV}$ (which can be
interesting for phenomenological reasons involving heavy vector-like
non-MSSM matter).
Finally, we note that case iii) reduces to case ii). To see this,
note, that the structure of \ensuremath{\mathcal{K}} and $\ensuremath{\mathcal{W}}$ given in case iii) does
not the change the arguments given for case i) which implies that in
case iii) we still find $\langle\phi_{(U)}\rangle=0$. This, however,
immediately gives us \begin{equation}
\left.\ensuremath{\mathcal{W}}\right|_{\langle\phi_{(U)}\rangle=0}\supset \lambda e^{-b T
}\left(\phi^{(T)}\right)^{N} \end{equation} which is case ii).
\subsection{Singlet directions which are $F$- and $D$-flat in global supersymmetry}
\label{subsubsec:FDflats}
There are many directions in singlet field space in our heterotic
constructions which are $F$- and $D$-flat in global supersymmetry.
Let us denote these fields by - $\phi_i^{(f)}$, and the remaining
set of non-flat directions in field space by $\chi_i$. $D$-flatness
entails that the $D$-terms do not depend on the $\phi_i^{(f)}$.
$F$-flatness implies that
$F_{\phi_i^{(f)}}=\partial_{\phi_i^{(f)}}\ensuremath{\mathcal{W}}(\phi_i^{(f)},\chi_i)=const.$
for all values of $\langle \phi_i^{(f)} \rangle$. Generically this
implies that $\langle \chi_i \rangle = 0$.
Simplifying to the case of a single $\chi$, this leads to a
consideration of 2 cases
\begin{itemize}
\item i)
\begin{eqnarray}
F_{\phi_i^{(f)}}&=&0\quad\forall\, \phi_i^{(f)}\nonumber\\
\Rightarrow\;\ensuremath{\mathcal{W}} & \supset& e^{- b T} \ \chi \
\mathfrak{f}(\phi_i)\quad\bigvee\quad\ensuremath{\mathcal{W}} \supset e^{- b T}
\chi^p \ \mathfrak{f}(\phi_i)\;,\; p \geq 2 \end{eqnarray}
\item ii)
\begin{eqnarray}
F_{\phi_i^{(f)}}&=&const.\neq0\quad\forall\, \phi_i^{(f)}\nonumber\\
\Rightarrow\quad\ensuremath{\mathcal{W}}&\supset& \lambda e^{-b T} f(\tilde
\phi_j)\phi_i^{(f)} \end{eqnarray}
\end{itemize}
where the $\tilde \phi_j$ VEVs are assumed fixed by other terms in
the superpotential and $\mathfrak{f}$ is an arbitrary function of
its argument.
We consider first case i). At the supersymmetric minimum satisfying
$\partial_\chi \ensuremath{\mathcal{W}} =
\partial_{\phi_i} \ensuremath{\mathcal{W}} = 0$, we have $\langle \chi \rangle = 0$ with $\langle
\phi_i \rangle$ arbitrary (subject, for the first case only, to the
condition $\mathfrak{f}(\phi_i) = 0$). In this example we have $
\chi \in \{ \chi_i \} $ and $\phi_i \in \{ \phi_i^{(f)} \} $. Note
the fields $\phi_i^{(f)}$ effectively do not appear in the
superpotential at its minimum.
We now argue that the fields $\phi_i^{(f)}$ are stabilized by the
corrections from supergravity in the $F$-term scalar potential.
Namely, consider for sake of simplicity the case of a single such
field $\phi^{(f)}$ and $\chi$ with
\begin{eqnarray} \ensuremath{\mathcal{K}}&=&-3\log(T+\bar T)+ c\overline\phi^{(f)}\phi^{(f)} + c' \overline\chi \chi\nonumber\\
\partial_\chi \ensuremath{\mathcal{W}} &=& \partial_{\phi^{(f)}} \ensuremath{\mathcal{W}} \equiv 0\quad {\rm for} \quad \langle \chi \rangle =
0\end{eqnarray} we get the F-term scalar potential in supergravity to be (for
the twisted sector case ii) we find the scalar potential to be \begin{eqnarray}
V_F&=&e^\ensuremath{\mathcal{K}}\left(\ensuremath{\mathcal{K}}^{\phi^{(f)}\bar\phi^{(f)}}|D_{\phi^{(f)}}\ensuremath{\mathcal{W}}|^2+\ensuremath{\mathcal{K}}^{\chi\bar\chi}|D_{\chi}\ensuremath{\mathcal{W}}|^2
+\ensuremath{\mathcal{K}}^{T\bar T}|D_T \ensuremath{\mathcal{W}}|^2-3|\ensuremath{\mathcal{W}}|^2\right)\nonumber\\
&=&e^\ensuremath{\mathcal{K}}\left(c\overline\phi^{(f)}\phi^{(f)} - \kappa \right)\cdot|\ensuremath{\mathcal{W}}|^2\nonumber\\
&\approx&|\ensuremath{\mathcal{W}}|^2\cdot\left[-c(\kappa
-1)\overline\phi^{(f)}\phi^{(f)}-\frac{c^2(\kappa -
2)}{2}\left(\overline\phi^{(f)}\phi^{(f)}\right)^2+\right. \nonumber \\
&&\left.-\frac{c^3(\kappa-3)}{6}\left(\overline\phi^{(f)}\phi^{(f)}\right)^3
+\frac{c^4(4-\kappa)}{24}\left(\overline\phi^{(f)}\phi^{(f)}\right)^4+\ldots\right]
\end{eqnarray} Note, we maintain $\langle \chi \rangle = 0$, $\ensuremath{\mathcal{W}} \neq 0$ is
due to other sectors of the theory and
$\kappa=(3-\ensuremath{\mathcal{K}}^{T\bar T}|D_T \ensuremath{\mathcal{W}}|^2/|\ensuremath{\mathcal{W}}|^2)\leq3$ is a
positive semi-definite number of order 3. This scalar potential is
unbounded from above at large field values, $\phi^{(f)}$, thus
driving the VEV to large-field value. To this order in $V_F$ we find
\begin{equation} \langle \phi^{(f)}\rangle\sim\frac{1}{\sqrt{c}}\quad. \end{equation} This
implies that supergravity effects will serve to stabilize all the
globally supersymmetric {\it and} $F$- and $D$-flat singlet fields
generically at large values of ${\cal O}(1)$. Note, that the
non-perturbative effects coming from gaugino-condensation in the
hidden sector will add dependence of $\ensuremath{\mathcal{W}}$ on $\phi^{(f)}$ beyond
the global mini-landscape analysis. This may render $\kappa$ a weak
function of $\phi^{f}$ such that we may for some of the globally
supersymmetric and $F$- and $D$-flat fields $\phi^{(f)}$ have
$\kappa<1$ at small $\phi^{(f)}$ while $1<\kappa<3$ at larger values
of $\phi^{(f)}$. In this situation the involved $\phi^{(f)}$-type
singlets will acquire vacua at both $\langle\phi^{(f)}\rangle=0$ and
$\langle\phi^{(f)}\rangle\sim1/\sqrt{c}$. The $\chi$-like fields
will have their VEVs near the origin, i.e. they may be shifted from
the origin by small SUSY breaking effects.
Let us now turn to case ii) of $F$-flat but {\it non}-supersymmetric
singlet directions and look for vacua stabilizing $\phi^{(f)}\ll 1$
using again \begin{equation} \ensuremath{\mathcal{K}}=-3\log(T+\bar T)+ \overline\phi^{(f)}\phi^{(f)}
+ \overline\chi \chi\quad. \end{equation} The scalar potential is \begin{eqnarray}
V_F&=& e^\ensuremath{\mathcal{K}}\left[\ensuremath{\mathcal{K}}^{T\bar T}\underbrace{\left\langle D_{T}\ensuremath{\mathcal{W}}\right\rangle}_{=F_T}\overline{\partial_T\ensuremath{\mathcal{W}}}+c.c.+\ensuremath{\mathcal{K}}^{\phi^{(f)}\bar\phi^{(f)}}|D_{\phi^{(f)}}\ensuremath{\mathcal{W}}|^2\right]\nonumber\\
&\sim&\left\{\ensuremath{\mathcal{K}}^{T\bar T}F_T\cdot b \lambda e^{-b T} f(\chi)\phi^{(f)}+c.c.\right.\nonumber\\
&&\left.+\ensuremath{\mathcal{K}}^{\phi^{(f)}\bar\phi^{(f)}}\left[ \lambda e^{-b T}
f(\chi)(1+\bar\phi^{(f)}\phi^{(f)})+\bar\phi^{(f)}\langle\ensuremath{\mathcal{W}}\rangle\right]^2\right\}\quad.
\end{eqnarray} In the desired regime of $\phi^{(f)}\ll 1$ this gives us two
sub-cases:
\begin{itemize}
\item iia)
\[\ensuremath{\mathcal{K}}^{\phi^{(f)}\bar\phi^{(f)}}F_{\phi^{(f)}}\ll \ensuremath{\mathcal{K}}^{T\bar T}F_T\]
\item iib)
\[\ensuremath{\mathcal{K}}^{\phi^{(f)}\bar\phi^{(f)}}F_{\phi^{(f)}}\gg \ensuremath{\mathcal{K}}^{T\bar T}F_T\]
\end{itemize}
In case iia) $\phi^{(f)}\ll1$ implies that
$F_{\phi^{(f)}}\equiv\lambda e^{-b T}
f(\langle\chi\rangle)\ll\langle\ensuremath{\mathcal{W}}\rangle$ and thus
$\partial_{\phi^{(f)}}V_F=0$ gives us \begin{equation}
\langle\phi^{(f)}\rangle\sim\frac{\langle
F_{\phi^{(f)}}\rangle}{\langle\ensuremath{\mathcal{W}}\rangle}\ll1 \end{equation} which is thus a
self-consistent vacuum.
In the opposite situation we get $F_{\phi^{(f)}}\equiv\lambda e^{-b
T} f(\langle\chi\rangle)\gg\langle\ensuremath{\mathcal{W}}\rangle,\langle F_T\rangle$.
Using again $\phi^{(f)}\ll1$ this leads to \begin{equation}
\langle\phi^{(f)}\rangle\sim\frac{\langle F_T\rangle}{\langle
F_{\phi^{(f)}}\rangle}\ll1\quad. \end{equation} Thus, even the $F$-flat but
non-supersymmetric singlet directions of case ii) get stabilized by
supersymmetry breaking effects from the bulk moduli stabilization at
generically small but non-zero VEVs.
This property, of all $F$- and $D$-flat singlet fields generically
acquiring non-zero VEVs from supersymmetry breaking in the bulk
moduli stabilizing sector through supergravity, dynamically ensures
the decoupling of all vector-like non-MSSM matter at low-energies as
checked in global supersymmetry for the mini-landscape setup.
Note, that the overall vacuum structure of the F-flat singlet fields
implicates a choice of initial conditions. The amount of non-MSSM
vector-like extra matter in the mini-landscape constructions which
decouples from low energies depends on the choice of the globally
$F$-flat singlets $\phi^{(f)}_i$ placed at their non-zero VEV vacuum
instead of their zero VEV vacuum. Thus, the choice of initial
conditions in the vacuum distribution among the set of globally
$F$-flat singlet fields characterizes how close to the MSSM one can
get when starting from one of the mini-landscape models.
Assuming now that one finds successful eternal inflation occurring
somewhere in the mini-landscape, this choice of initial conditions
turns into a question of cosmological dynamics. In this situation,
all possible initial conditions of the set of globally $F$-flat
singlets were potentially realized in a larger multiverse. The
choice of initial conditions on the singlets in the globally
$F$-flat sector would then be amenable to anthropic arguments and
might be eventually determined by selection effects.
\section{SUSY spectrum}
\label{sec:spectrum}
Now that we understand how SUSY is broken, we can calculate the
spectrum of soft masses. The messenger of SUSY breaking is mostly
gravity, however, there are other contributions from gauge and
anomaly mediation.
\subsection{Contributions to the soft terms}
At tree level, the general soft terms for gravity mediation are
given in References
\cite{Brignole:1993dj,Brignole:1995fb,Kawamura:1996wn,Kawamura:1996bd,
Brignole:1997dp}. The models described in this paper contain an
additional contribution from the $F$-term of a scalar field
$\phi_2$. Following References
\cite{Brignole:1993dj,Brignole:1995fb, Brignole:1997dp}, we define
\begin{equation} \label{F_terms_defn}
F^I \equiv e^{\ensuremath{\mathcal{K}}/2}\ensuremath{\mathcal{K}}^{I\bar{J}}\left(\bar{\ensuremath{\mathcal{W}}}_{\bar{J}} + \bar{\ensuremath{\mathcal{W}}}\ensuremath{\mathcal{K}}_{\bar{J}}\right).
\end{equation}
\subsubsection{SUGRA effects}
{\em\textit{Gaugino masses}}
The tree level gaugino masses are given by
\begin{equation}
M_a^{(0)} = \frac{g_a^2}{2}F^n\partial_n f_a(S) = \frac{g_a^2}{2}F^S.
\end{equation}
At tree level, the gauge kinetic function in heterotic string theory is linear
in the dilaton superfield $S$, and only dependent on the $T$ modulus at one loop.
It is important to note the enhancement of $F^S$ relative to $F_S$: naively, one
might guess that loop corrections to the gaugino masses might be important,
however
\begin{equation}
F^S >> \frac{F^T}{16\pi^2},
\end{equation}
thus loop corrections will be neglected. \vspace{.1in}
\noindent{\em\textit{A Terms}}
At tree level, the $A$ terms are given by
\begin{equation}
A_{IJK}^{(0)} = F^n \partial_n \ensuremath{\mathcal{K}} + F^n\partial_n \log\frac{\ensuremath{\mathcal{W}}_{IJK}}{\kappa_I \kappa_J \kappa_K},
\end{equation}
where
\begin{equation}
\ensuremath{\mathcal{W}}_{IJK} \equiv \frac{\partial^3 \ensuremath{\mathcal{W}}}{\partial \Phi^I\partial\Phi^J\partial\Phi^K}
\end{equation} and $\ensuremath{\mathcal{K}}$ is the K\"{a}hler potential. Neglecting $U$ dependence, we have
\begin{equation}
\ensuremath{\mathcal{K}} \supset \Phi_I\bar{\Phi}^I \prod_i \left(T_i+\bar{T}_i\right)^{-n^i_I} \Rightarrow \kappa_I \equiv \prod_i \left(T_i+\bar{T}_i\right)^{-n^i_I}.
\end{equation} The $\kappa_I$ are the K\"{a}hler metrics for the chiral multiplets,
$\Phi_I$, where as the $A$ terms are expressed in terms of
canonically normalized fields. As before, the modular weights of
the matter field are given by $n^i_I$.
In general, there are also tree level contributions to $A$ terms
proportional to \begin{equation} - \frac{F_{\phi_2}}{\langle \phi_2 \rangle}
\frac{\partial \log \ensuremath{\mathcal{W}}_{IJK}}{\partial \log \phi_2} . \end{equation} These
terms may be dominant, but unfortunately they are highly model
dependent. They may give a significant contribution to $A_b$ and
$A_\tau$, but in fact we find that the details of the low energy
spectrum are not significantly effected. \vspace{.1in}
\noindent{\em\textit{Scalar masses}}
The tree level scalar masses are given by
\begin{equation}
\left(M_I^{(0)}\right)^2 = m_{3/2}^2 - F^n\bar{F}^{\bar{m}}\partial_n \partial_{\bar{m}} \log \kappa_I +
g_G^2 \ f \ q^I_A \ \langle D_A \rangle \ \kappa_I ,
\end{equation}
where $g_G^2 = 1/\re{S_0}$ and we have implicitly assumed that the
K\"{a}hler metric is diagonal in the matter fields. The factor $f$
re-scales the $U(1)_A$ charges $q_A$ from the mini-landscape
``benchmark" model 1 \cite{Lebedev:2007hv}, so they are consistent
with the charges $q_A'$ in our mini-version of the mini-landscape
model. We have $q_A' = q_A \ f = q_A \ \frac{48 \pi^2}{Tr Q}
\delta_{GS}$ with $\delta_{GS} =\frac{N_f}{4 \pi^2}$ (Eqn.
\ref{eq:GS}) and $Tr Q = \frac{296}{3}$ (Eqn. E.5,
\cite{Lebedev:2007hv}) such that $\frac{Tr(q')}{4 \pi^2} =
\delta_{GS}$.
Again neglecting $U$ dependence, the K\"{a}hler metric for the
matter fields depends only on the $T$ moduli, and we find
\begin{equation} \label{tree_level_scalars}
\left(M_I^{(0)}\right)^2 = m_{3/2}^2 - \sum_i \frac{n^i_I \left|F^{T_i}\right|^2}{\left(T_i+\bar{T}_i\right)^2} +
g_G^2 \ f \ q^I_A \ \langle D_A \rangle / (2 \re{T_0})^{n^3_I} .
\end{equation}
\vspace{.1in}
\noindent{\em $\mu$ and $B\mu$ terms}
The $\mu$ term can come from two different sources:
\begin{equation}
\ensuremath{\mathcal{K}} \supset Z(T_i+\bar{T}_i,U_j+\bar{U}_j,...) \mathbf{H}^u \mathbf{H}^d, \,\,\, \ensuremath{\mathcal{W}} \supset
\tilde \mu(\mathbf{s}_I,T_i,U_j,...) \mathbf{H}^u \mathbf{H}^d.
\end{equation}
In the orbifold models, K\"{a}hler corrections have not been
computed, so the function $Z$ is \textit{a priori} unknown. Such a
term could contribute to the Giudice-Masiero mechanism
\cite{Giudice:1988yz}. When both $\tilde \mu$ and $Z$ vanish, the
SUGRA contribution to the $\mu/B\mu$ terms vanish. On the other
hand, in the class of models which we consider, we know that vacuum
configurations exist such that $\tilde \mu = 0$ to a very high order
in singlet fields. Moreover $\tilde \mu \propto \langle \ensuremath{\mathcal{W}} \rangle$
which vanishes in the supersymmetric limit, but obtains a value
$w_0$ at higher order in powers of chiral singlets. If $\mu$ is
generated in this way, there is also likely to be a Peccei-Quinn
axion \cite{Kim:1983dt,Kim:1994eu}. Finally, supergravity effects
will also generate a $B \mu$ term. \vspace{.1in}
\noindent{\em Loop corrections}
Finally, one can consider loop corrections to the tree level
expressions in
\cite{Brignole:1993dj,Brignole:1995fb,Brignole:1997dp}. This was
done in References \cite{Binetruy:2000md, Binetruy:2003ad}, where
the complete structure of the soft terms (at one loop) for a generic
(heterotic) string model were computed in the effective supergravity
limit. We have applied the results of \cite{Binetruy:2000md,
Binetruy:2003ad} to our models and find, at most, around a 10\%
correction to the tree level results of
\cite{Brignole:1993dj,Brignole:1995fb,Brignole:1997dp}.\footnote{In
estimating this result, we have assumed that the mass terms of the
Pauli-Villars fields do not depend on the SUSY breaking singlet
field $\phi_2$, and that the modular weights of the Pauli-Villars
fields obey specific properties.}
\subsubsection{Gauge mediation}
The ``mini-landscape" models generically contain vector-like exotics
in the spectrum. Moreover it was shown that such states were
necessary for gauge coupling unification \cite{Dundee:2008ts}. The
vector-like exotics obtain mass in the supersymmetric limit by
coupling to scalar moduli, thus they may couple to the SUSY
breaking field $\phi_2$. We will consider the following light
exotics to have couplings linear in the field $\phi_2$:
\begin{equation} \label{light_exotica}
n_3\times (\mathbf{3},1)_{1/3} + n_2 \times (1,\mathbf{2})_0 + n_1 \times (1,1)_{-1} + \text{h.c.}
\end{equation}
where the constants $n_i$ denote the multiplicity of states and (see
Table 7 of Reference \cite{Dundee:2008ts})
\begin{equation}
n_3 \leq 4\text{~and~}n_2\leq 3\text{~and~}n_1 \leq 7.
\end{equation}
The gauge mediated contributions split the gaugino masses by an amount proportional
to the gauge coupling:
\begin{eqnarray}
M_3^{(1)}|_{\text{gmsb}} &=& n_3 \frac{g_3^2}{16\pi^2}\frac{F^{\phi_2}}{\vev{\phi_2}}, \\
M_2^{(1)}|_{\text{gmsb}} &=& n_2 \frac{g_2^2}{16\pi^2}\frac{F^{\phi_2}}{\vev{\phi_2}}, \\
M_1^{(1)}|_{\text{gmsb}} &=& \frac{n_3+3n_1}{10} \frac{g_1^2}{16\pi^2}\frac{F^{\phi_2}}{\vev{\phi_2}}.
\end{eqnarray}
It is interesting to note that this becomes more important as
$\vev{\phi_2}$ decreases/$F^{\phi_2}$ increases, or if there are a
large number of exotics present.
The scalar masses in gauge mediation come in at two loops, and receive corrections
proportional to
\begin{equation}
\left(M_I\right)^2|_{\text{gmsb}} \sim \left(\frac{1}{16\pi^2}\right)^2\left(\frac{F_{\phi_2}}{\phi_2}\right)^2.
\end{equation}
Unlike in the case of the gaugino masses, however, the tree level
scalar masses are set by the gravitino mass. Typically
\begin{equation}
16\pi^2 m_{3/2} >> \frac{F_{\phi_2}}{\phi_2},
\end{equation}
and the gauge mediation contribution gives about a 10\% correction to the scalar masses,
in our case. We will neglect their contributions in the calculation of the soft masses below.
\subsection{Calculation of the soft terms -- relevant details from the ``mini-landscape"}
\label{sec:details}
Given the relative sizes of the $F$-terms in the
SUSY breaking sectors described in this paper, it is very difficult
to make model-independent statements. This stems from the fact that
$F^T$ plays a dominant role in the SUSY breaking. Because the
K\"{a}hler metrics for the matter fields have generally different
dependences on the $T$ modulus, the dependence of the soft terms on
$F^T$ is typically non-universal. Moreover, the couplings of the
SUSY breaking singlet field $\phi_2$ will necessarily depend on the
details of a specific model. Thus, in order to make \textit{any}
statements about the phenomenology of these models, we will have to
make some assumptions. With the general features of the
``mini-landscape'' models in mind, we will make the following
assumptions:
\begin{enumerate}
\item SUSY breaking is dominated by $F_{\phi_2} \neq 0$, $F_{T_3} \neq 0$, $F_S\neq 0$.
All other $F$ terms, including those due to the other $T$ and $U$ moduli, are subdominant;
\item \label{brane_exotics_oinly} the massless spectrum below $M_{\textsc{s}}$ contains some vector-like
exotics;
\item the untwisted sector contains the following Higgs and (3rd generation) matter multiplets:
$\mathbf{H}_u, \mathbf{H}_d, \mathbf{Q}_3, \mathbf{U}^c_3,
\mathbf{E}^c_3$;
\item the first two families have the same modular weights, see Table
\ref{tab:modular_weights_MSSM};
\item the SUSY breaking field, $\phi_2$, lives in the untwisted, or second or fourth twisted sector, with a modular weight given by $n^3 = 0$; and
\item we neglect possible $\phi_2$ dependence of the effective
Yukawa terms.
\end{enumerate}
Let us examine these assumptions in some more detail.
In general, gauge coupling unification in the ``mini-landscape"
models seems to require the existence of light vector-like exotics
\cite{Dundee:2008ts}, whose masses can be as small as
$\order{10^{9}{\rm ~GeV~}}$. We further assume that these exotics couple
to the SUSY breaking field $\phi_2$, giving a gauge mediated
contribution to the gaugino masses above. We will make this
contribution to the soft terms explicit in what follows. In
assumption \ref{brane_exotics_oinly} we have specialized to the case
where only ``brane-localized'' exotics are present in the model.
These are states which come from the first and third twisted sectors
of the model, and we refer the reader to \cite{Lebedev:2007hv,
Dundee:2008ts} for more details.
The top quarks and the up Higgses live in the bulk and the string
selection rules allow for the following coupling in the
superpotential:
\begin{equation} \label{third_family_top}
\ensuremath{\mathcal{W}} \supset c\, \mathbf{Q}_3 \mathbf{H}_u \mathbf{U}^c_3.
\end{equation}
The coupling $c$ is a pure number of \order{1}, and is free of any
dependence on the moduli. The down and lepton Yukawas are a bit
more involved, as they arise at a higher order in the stringy
superpotential. We will take them to be of the following form:
\begin{equation} \label{third_family_bottom}
\ensuremath{\mathcal{W}} \supset \eta(T_1)^{p_1}\eta(T_2)^{p_2}\eta(T_3)^{p_3} \
\left(f_1(\vev{\mathbf{s}_I^5})\mathbf{Q}_3 \mathbf{H}_d \mathbf{D}^c_3 + f_2(\vev{\mathbf{s}_I^5})\mathbf{L}_3 \mathbf{H}_d \mathbf{E}^c_3\right).
\end{equation}
The $\mathbf{s}_I$ are other singlet fields in the model (excluding
the SUSY breaking singlet field, $\phi_2$, as per our assumptions),
and the numbers $p_1,p_2$ and $p_3$ are calculable in principle,
given knowledge of the modular weights of the $\mathbf{s}_I$. As one
might expect, the expressions for the $A$ terms explicitly depend on
the value of $p_3$ in such a way that changing its value may result
in a significant change in $A_b$ and $A_{\tau}$ at the string scale.
The impact on the weak scale observables is much less severe,
however, giving a correction of a few percent to the gaugino masses,
and leaving the squark and slepton masses virtually unchanged.
Motivated by the modular weight assignments in Table
\ref{tab:modular_weights_MSSM}, we will choose $p_3 = 0$. Note this
choice gives us universal $A$ terms for the third generation.
One of the nice features of the ``mini-landscape" models is the
incorporation of a discrete ($D_4$) symmetry between the first two
families in the low energy effective field theory. Because of this
symmetry, we expect the modular weights of these matter states to be
the same \cite{Ko:2007dz}, see Table \ref{tab:modular_weights_MSSM}.
This will turn out to be very beneficial in alleviating the flavor
problems that are generic in gravity mediated models of SUSY
breaking: the scalar masses (at tree level) are given by a universal
contribution (the gravitino mass squared) plus a contribution
proportional to the modular weight. If the modular weights are the
same between the first two generations, then the leading order
prediction is for degenerate squark and slepton masses in the two
light generations. Other contributions to the scalar masses come
from gauge mediation and anomaly mediation, which do not introduce
any new flavor problems into the low energy physics.
\begin{table}
\centering
\begin{tabular}{|c|c|cc|}
\hline
MSSM particle&Modular Weight $\vec{n}$& \multicolumn{2}{c|}{$U(1)_A$ charge}\\
\hline $\mathbf{Q}_3$&$(0,1,0)$ & \multicolumn{2}{c|}{$4/3$} \\
$\mathbf{U}^c_3$&$(1,0,0)$ & \multicolumn{2}{c|}{$2/3$} \\
$\mathbf{D}^c_3$&$\left(\frac{1}{3},\frac{2}{3},0\right)$ & \multicolumn{2}{c|}{$8/9$} \\
$\mathbf{L}_3$&$\left(\frac{2}{3},\frac{1}{3},0\right)$ & \multicolumn{2}{c|}{$4/9$} \\
$\mathbf{E}^c_3$ & $(1,0,0)$ & \multicolumn{2}{c|}{$2/3$} \\
\multirow{2}{*}{first two gen.}&\multirow{2}{*}{$\left(\frac{5}{6},\frac{2}{3},\frac{1}{2}\right)$}
& $7/18$ & (${\bf 10}$) \\
&& $-5/18$ & (${\bf \bar 5}$) \\
$\mathbf{H}_u$&$(0,0,1)$ & \multicolumn{2}{c|}{$-2$} \\
$\mathbf{H}_d$&$(0,0,1)$ & \multicolumn{2}{c|}{$+2$} \\
\hline \hline \end{tabular}
\caption{\label{tab:modular_weights_MSSM}Modular weights of the MSSM
states in the ``mini-landscape" benchmark model 1A. For the first
two generations, the $U(1)_A$ charges differ depending on whether
the particle is in the {\bf 10} or ${\bf \bar 5}$ of $SU(5)$. See
\cite{Lebedev:2007hv} for details.} \end{table}
\subsection{Hierarchy of $F$-terms}
Note, in Section \ref{sec:one_gaugino_condensate}, we find (roughly)
\begin{equation}
F_T >> F_S \gtrsim F_{\phi_2},
\end{equation} for Cases 1, 2 and 3; and
\begin{equation}
F_T \gtrsim F_{\phi_2} >> F_S,
\end{equation} for Cases 4 and 5, where
\begin{equation}
F_I \equiv \ensuremath{\mathcal{W}}_I + \ensuremath{\mathcal{W}} \ensuremath{\mathcal{K}}_I.
\end{equation}
When one includes the relevant factors of the K\"{a}hler metric, we
have (Table \ref{tab:F_terms})
\begin{equation}
F^T > F^S >> F^{\phi_2}
\end{equation} for Cases 1, 2 and 3; and
\begin{equation}
F^T >> F^S \sim F^{\phi_2}
\end{equation} for Cases 4 and 5.
$F^S$ is enhanced by a factor of $\ensuremath{\mathcal{K}}^{S\bar{S}}\sim (2+2)^2$, while
$F^{\phi_2}$ is decreased by a factor of
$\ensuremath{\mathcal{K}}^{\phi_2\bar{\phi}_2}\sim(2)^{-1/2}$.\footnote{This is due to
the assumed modular weight of the field $\phi_2$ (assumption 5 in
Section \ref{sec:details}).} This means that although the singlet
field $\phi_2$ was a dominant source of SUSY breaking, it is the
least important when computing the soft terms, given the one
condensate hidden sector of the known ``mini-landscape" models
studied in Section \ref{sec:one_gaugino_condensate}.\footnote{In
racetrack models $F_S$ is suppressed by more than an order of
magnitude. In these cases $F_{\phi_2}$ is dominant
\cite{Gallego:2008sv}.} Taking the details of the
``mini-landscape" models into account, the soft terms at the string
scale are given in Table \ref{tab:string_scale_BCs}.
\begin{table}[t!]
\renewcommand\arraystretch{1.2}
\centering
\noindent\makebox[\textwidth]{ \scalebox{0.75}{
\begin{tabular}{|c|c|c|c|c|c|}
\hline
& Case 1 & Case 2 & Case 3 & Case 4 & Case 5 \\
\hline
$F^S$ & $6.6\times10^{-16}$ & $3.7\times10^{-16}$ & $4.2\times10^{-16}$ & $2.7\times10^{-16}$ & $2.1\times10^{-16}$\\
\hline
$F^T$ & $-2.2\times10^{-15}$ & $-1.2\times10^{-15}$ & $-1.4\times10^{-15}$ & $1.6\times10^{-15}$ & $2.2\times10^{-15}$\\
\hline
$F^{\phi_2}$ & $-1.1\times10^{-17}$ & $-6.5\times10^{-18}$ & $-7.7\times10^{-18}$ & $1.9\times10^{-16}$ & $1.8\times10^{-16}$\\
\hline
\end{tabular}
}} \caption{\label{tab:F_terms} The hierarchy of $F$ terms in the
five examples of the single condensate model we studied. Note that
$F^{\Phi}$ is defined in Eqn. (\ref{F_terms_defn}). All of the $F$
terms contribute to the soft masses, as they are all within an order
of magnitude.}
\end{table}
In the five chosen Cases, 2, 3 and 4 have a gravitino mass less than
2 TeV. The value of the gravitino mass can be adjusted by varying
$w_0$. For Cases, 1, 3 (4) the Higgs up (down) mass squared is
negative. This is a direct result of the sign of $D_A$ and the
$U(1)_A$ charge of the Higgs' (see Table
\ref{tab:modular_weights_MSSM} for the $U(1)_A$ charges of all the
MSSM states).\footnote{Note, it is well known that the $D$-term VEV
in supergravity is of order $\langle F^i \rangle^2$
\cite{Kawamura:1995ec,Kawamura:1996bd}. It is given by the relation
\begin{equation} \langle D_A \rangle = 2 M_A^{-2} \langle F^i \rangle \langle
F_j^* \rangle \langle \partial_i \partial^j D_A \rangle . \end{equation} Thus
the $D$-term contribution to the vacuum energy is negligible, but
its contribution to scalar masses can be significant. Since $|F^S|^2
< |F^T|^2$, $F^T$ is dominant in the above relation. However, the
K\"{a}hler metric of $\phi_2$ which spontaneously breaks $U(1)_A$,
in our case, does not include $T$, i.e. $ \langle (\partial_T
\partial^T D_A) \rangle = 0$. Hence $\langle D_A \rangle$ is
suppressed compared with $|F^T|^2/M_{Pl}^2$, i.e. $\langle D_A
\rangle : |F^T|^2/M_{Pl}^2 = |F^S|^2 : |F^T|^2$ where we used
$\langle (\partial_S \partial^S D_A)\rangle = (M_A/M_{Pl})^2$,
because of the S-dependent FI term. We thank T.~Kobayashi, private
communication, for this analysis. However, it should be clear that
we have also used the freedom available in the Coleman-Weinberg
one-loop correction to further adjust the value of the $D$-term.}
Note, the first and second generation squarks and sleptons are
lighter than the third generation states at the string scale. This
is a consequence of the significant $T$ modulus contribution to the
first and second generation squark and slepton masses, due to their
modular weights, Table \ref{tab:modular_weights_MSSM}. Finally we
have included the possible gauge mediated SUSY breaking contribution
to the gaugino masses, Table \ref{tab:string_scale_BCs}. This
contribution is only significant for Cases 4 and 5, due to the
larger value of $F_{\phi_2}$ in these cases.
\begin{table}[t!]
\renewcommand\arraystretch{1.2}
\centering \noindent\makebox[\textwidth]{ \scalebox{0.75}{
\begin{tabular}{|c|cc|cc|cc|cc|cc|}
\cline{2-11}
\multicolumn{1}{c|}{ }&\multicolumn{10}{c|}{\textit{All Masses in GeV}} \\
\hline
Parameter&\multicolumn{2}{c|}{Case 1}&\multicolumn{2}{c|}{Case 2}&\multicolumn{2}{c|}{Case 3}&\multicolumn{2}{c|}{Case 4}&\multicolumn{2}{c|}{Case 5}\\
\hline
$m_{3/2}$&\multicolumn{2}{c|}{2159}&\multicolumn{2}{c|}{1350}&\multicolumn{2}{c|}{1133}&\multicolumn{2}{c|}{1808}&\multicolumn{2}{c|}{2375}\\
\hline
$m_{H_u}$&\multicolumn{2}{c|}{$478i$}&\multicolumn{2}{c|}{168}&\multicolumn{2}{c|}{$372i$}&\multicolumn{2}{c|}{688}&\multicolumn{2}{c|}{384}\\
$m_{H_d}$&\multicolumn{2}{c|}{679}&\multicolumn{2}{c|}{216}&\multicolumn{2}{c|}{495}&\multicolumn{2}{c|}{$476i$}&\multicolumn{2}{c|}{251}\\
\hline
$M_1$&\multicolumn{2}{c|}{$362-0.3n_1-0.1n_3$}&\multicolumn{2}{c|}{$206-0.2n_1-0.1n_3$}&\multicolumn{2}{c|}{$243-0.2n_1-0.1n_3$}&\multicolumn{2}{c|}{$158+13n_1+4n_3$}&\multicolumn{2}{c|}{$118+7n_1+2n_3$}\\
$M_2$&\multicolumn{2}{c|}{$362-1n_2$}&\multicolumn{2}{c|}{$206+1n_2$}&\multicolumn{2}{c|}{$243-1n_2$}&\multicolumn{2}{c|}{$158+45n_2$}&\multicolumn{2}{c|}{$118+23n_2$}\\
$M_3$&\multicolumn{2}{c|}{$362-1n_3$}&\multicolumn{2}{c|}{$206+1n_3$}&\multicolumn{2}{c|}{$243-1n_3$}&\multicolumn{2}{c|}{$158+45n_3$}&\multicolumn{2}{c|}{$118+23n_3$}\\
\hline
$A_t$&\multicolumn{2}{c|}{3901}&\multicolumn{2}{c|}{2466}&\multicolumn{2}{c|}{1974}&\multicolumn{2}{c|}{-3690}&\multicolumn{2}{c|}{-4798}\\
$A_{b}$&\multicolumn{2}{c|}{3901}&\multicolumn{2}{c|}{2466}&\multicolumn{2}{c|}{1974}&\multicolumn{2}{c|}{-3690}&\multicolumn{2}{c|}{-4798}\\
$A_{\tau}$&\multicolumn{2}{c|}{3901}&\multicolumn{2}{c|}{2466}&\multicolumn{2}{c|}{1974}&\multicolumn{2}{c|}{-3690}&\multicolumn{2}{c|}{-4798}\\
\hline
&Gen.~1,2&Gen.~3&Gen.~1,2&Gen.~3&Gen.~1,2&Gen.~3&Gen.~1,2&Gen.~3&Gen.~1,2&Gen.~3\\
\cline{2-11}
$m_{\tilde{q}}$ & 1580 & 2288 & 966 & 1355 & 895 & 1446 & 1262 & 1657 & 1691 & 2361 \\
$m_{\tilde{u}^c}$ & 1580 & 2225 & 966 & 1353 & 895 & 1299 & 1262 & 1734 & 1691 & 2368 \\
$m_{\tilde{d}^c}$ & 1521 & 2246 & 964 & 1354 & 757 & 1350 & 1330 & 1709 & 1697 & 2366 \\
$m_{\tilde{\ell}}$ & 1580 & 2203 & 964 & 1352 & 757 & 1246 & 1330 & 1759 & 1697 & 2370 \\
$m_{\tilde{e}^c}$ & 1580 & 2225 & 966 & 1353 & 895 & 1299 & 1262 & 1734 & 1691 & 2368 \\
\hline
\hline
\end{tabular}
} } \caption{\label{tab:string_scale_BCs}Boundary conditions at the
string scale. $n_3,n_2,n_1$ refer to possible intermediate mass
vector-like exotics which couple to the SUSY breaking field
$\phi_2$, see Eqn. (\ref{light_exotica}).}
\end{table}
\subsection{Weak scale observables}
We do not intend this work to be a comprehensive study of the
parameter space of these models, so we will limit our weak scale
analysis to the five cases studied in the single condensate model
presented in this paper. The points are chosen subject to the
following constraints:
\begin{itemize}
\item $m_{h^0}\Big|_{\text{LEP}} \gtrsim 114.4 {\rm ~GeV~}$,
\item successful electroweak symmetry breaking,
\item $m_{\tilde{\chi}^{\pm}} \gtrsim 94{\rm ~GeV~}$, and
\item the low energy spectrum is free of tachyons.
\end{itemize}
Note that we take $\text{sgn}(\mu) > 0$ and vary $\tan\beta$, and
the number, $n_i$, of ``messenger'' exotics. We stay in the region
of small to moderate $\tan\beta$ as the ``mini-landscape" models do
not tend to predict unification of the third family Yukawas. This
can be seen from Eqns. (\ref{third_family_top}) and
(\ref{third_family_bottom}), for example.
Using {\tt SoftSUSY} (v3.1) \cite{Allanach:2001kg}, we preformed the RGE running
from the string scale to the weak scale. We use the current value of the top quark
mass \cite{tev_EW_working_group}
\begin{equation}
m_{top}\Big|_{\text{world~avg.}} = 173.1 {\rm ~GeV~}
\end{equation}
and the strong coupling constant at $\left(M_Z\right)$ \cite{Amsler:2008zzb}
\begin{equation}
\alpha_s(\left(M_Z\right)) = 0.1176.
\end{equation}
The $\mu$ parameter is obtained under the requirement of radiative electroweak
symmetry breaking, and is of order the gravitino mass, as expected. This implies
a fine tuning of order
\begin{equation}
\frac{\left(M_Z\right)^2}{m_{3/2}^2}\sim \order{10^{-2}} \text{~to~} \order{10^{-4}}.
\end{equation}
The results obtained from \texttt{SoftSUSY} are presented in Table
\ref{tab:weak_obs}. In this analysis, we have not included any
possible gauge mediated SUSY breaking contributions. This assumes
that all the vector-like exotics have mass at the string scale. In
Case 2 and 3 we have the smallest gravitino masses, so the lightest
SUSY partners. $\tan\beta = 25$ in order for the light Higgs mass
to be above the LEP bound. Note we assume a $\pm 2$ GeV theoretical
uncertainty in the Higgs mass. In all 5 cases the Higgs mass is
between the LEP bound and 121 GeV. All other Higgs masses are of
order the gravitino mass. In all 5 cases the gluino mass is less
than 1 TeV and of order 600 GeV or less in Cases 2, .. , 5. Thus
the gluino is very observable at the LHC. In all cases, the
lightest MSSM particle is the lightest neutralino. The
next-to-lightest neutralino and the lightest chargino are
approximately degenerate with mass of order twice the lightest
neutralino mass. In Cases 2, 3 and 4 the lightest stop has mass
less than 1 TeV. In Cases 2 and 4, the lightest stop is also the
lightest squark. Thus in these cases the gluino will predominantly
decay into a top - anti-top pair with missing energy (and possibly
two energetic leptons). In Case 3, the lightest down squarks of
the first two families are lighter than the lightest stop. In
these cases gluinos will decay significantly into two light quark
jets plus missing energy (and possibly two energetic leptons).
\begin{table}[t!]
\renewcommand\arraystretch{1.2}
\begin{center}
\noindent\makebox[\textwidth]{ \scalebox{0.75}{
\begin{tabular}{c|c|c|c|c|c|c|c|c|c|c|c|}
\cline{3-12}
\multicolumn{2}{c|}{ }&\multicolumn{10}{c|}{\textit{All Masses in GeV (defined at $M_{\textsc{w}} \approx 80$ GeV, unless otherwise noted.) }} \\
\hline
&Observable&\multicolumn{2}{c|}{Case 1}&\multicolumn{2}{c|}{Case 2}&\multicolumn{2}{c|}{Case 3}&\multicolumn{2}{c|}{Case 4}&\multicolumn{2}{c|}{Case 5}\\
\hline
\multirow{4}{*}{\rotatebox{90}{Inputs}}&$m_{3/2}$&\multicolumn{2}{c|}{2159}&\multicolumn{2}{c|}{1350}&\multicolumn{2}{c|}{1133}&\multicolumn{2}{c|}{1808}&\multicolumn{2}{c|}{2375}\\
&$\tan\beta$&\multicolumn{2}{c|}{10}&\multicolumn{2}{c|}{25}&\multicolumn{2}{c|}{25}&\multicolumn{2}{c|}{10}&\multicolumn{2}{c|}{4}\\
&$\text{sgn}(\mu)$&\multicolumn{2}{c|}{$+$}&\multicolumn{2}{c|}{$+$}&\multicolumn{2}{c|}{$+$}&\multicolumn{2}{c|}{$+$}&\multicolumn{2}{c|}{$+$}\\
&$n_1,n_2,n_3$&\multicolumn{2}{c|}{0,0,0}&\multicolumn{2}{c|}{0,0,0}&\multicolumn{2}{c|}{0,0,0}&\multicolumn{2}{c|}{0,0,0}&\multicolumn{2}{c|}{0,0,0}\\
\hline
\multirow{5}{*}{\rotatebox{90}{EWSB}}&$\mu({M_{\textsc{susy}}})$&\multicolumn{2}{c|}{2221}&\multicolumn{2}{c|}{1317}&\multicolumn{2}{c|}{1342}&\multicolumn{2}{c|}{1848}&\multicolumn{2}{c|}{2636}\\
&$m_{h^0}$&\multicolumn{2}{c|}{115.8}&\multicolumn{2}{c|}{113.3}&\multicolumn{2}{c|}{113.5}&\multicolumn{2}{c|}{121.4}&\multicolumn{2}{c|}{116.7}\\
&$m_{H^0}$&\multicolumn{2}{c|}{2299}&\multicolumn{2}{c|}{1161}&\multicolumn{2}{c|}{1368}&\multicolumn{2}{c|}{1731}&\multicolumn{2}{c|}{2717}\\
&$m_{A^0}$&\multicolumn{2}{c|}{2303}&\multicolumn{2}{c|}{1173}&\multicolumn{2}{c|}{1376}&\multicolumn{2}{c|}{1728}&\multicolumn{2}{c|}{2715}\\
&$m_{H^+}$&\multicolumn{2}{c|}{2305}&\multicolumn{2}{c|}{1176}&\multicolumn{2}{c|}{1379}&\multicolumn{2}{c|}{1730}&\multicolumn{2}{c|}{2716}\\
\hline
\multirow{3}{*}{\rotatebox{90}{$M_a({M_{\textsc{susy}}})$}} &$M_1$&\multicolumn{2}{c|}{151}&\multicolumn{2}{c|}{83}&\multicolumn{2}{c|}{100}&\multicolumn{2}{c|}{68}&\multicolumn{2}{c|}{53}\\
&$M_2$&\multicolumn{2}{c|}{277}&\multicolumn{2}{c|}{155}&\multicolumn{2}{c|}{185}&\multicolumn{2}{c|}{128}&\multicolumn{2}{c|}{100}\\
&$M_3$&\multicolumn{2}{c|}{773}&\multicolumn{2}{c|}{457}&\multicolumn{2}{c|}{538}&\multicolumn{2}{c|}{370}&\multicolumn{2}{c|}{279}\\
\hline
\rotatebox{90}{$\tilde{g}$}&$m_{\tilde{g}}$&\multicolumn{2}{c|}{914}&\multicolumn{2}{c|}{545}&\multicolumn{2}{c|}{630}&\multicolumn{2}{c|}{456}&\multicolumn{2}{c|}{365}\\
\hline
\multirow{6}{*}{\rotatebox{90}{Neut./Charg.}}&$m_{\tilde{\chi}_1^0}$&\multicolumn{2}{c|}{150}&\multicolumn{2}{c|}{83}&\multicolumn{2}{c|}{99}&\multicolumn{2}{c|}{68}&\multicolumn{2}{c|}{52}\\
&$m_{\tilde{\chi}_2^0}$&\multicolumn{2}{c|}{293}&\multicolumn{2}{c|}{164}&\multicolumn{2}{c|}{194}&\multicolumn{2}{c|}{136}&\multicolumn{2}{c|}{104}\\
&$m_{\tilde{\chi}_3^0}$&\multicolumn{2}{c|}{-2204}&\multicolumn{2}{c|}{-1306}&\multicolumn{2}{c|}{-1334}&\multicolumn{2}{c|}{-1835}&\multicolumn{2}{c|}{-2616}\\
&$m_{\tilde{\chi}_4^0}$&\multicolumn{2}{c|}{2206}&\multicolumn{2}{c|}{1307}&\multicolumn{2}{c|}{1335}&\multicolumn{2}{c|}{1836}&\multicolumn{2}{c|}{2617}\\ \cline{2-12}
&$m_{\tilde{\chi}_1^\pm}$&\multicolumn{2}{c|}{293}&\multicolumn{2}{c|}{164}&\multicolumn{2}{c|}{194}&\multicolumn{2}{c|}{136}&\multicolumn{2}{c|}{104}\\
&$m_{\tilde{\chi}_2^\pm}$&\multicolumn{2}{c|}{2214}&\multicolumn{2}{c|}{1313}&\multicolumn{2}{c|}{1341}&\multicolumn{2}{c|}{1839}&\multicolumn{2}{c|}{2622}\\
\hline
\multirow{8}{*}{\rotatebox{90}{Squarks/Sleptons}}&& Gen.~1,2&Gen.~3& Gen.~1,2&Gen.~3& Gen.~1,2&Gen.~3& Gen.~1,2&Gen.~3& Gen.~1,2&Gen.~3\\ \cline{3-12}
&$m_{\tilde{u}_1}$ & 1712 & 1542 & 1040 & 921 & 1013 & 987 & 1283 & 717 & 1677 & 1107 \\
&$m_{\tilde{u}_2}$ & 1704 & 2042 & 1038 & 1164 & 1006 & 1336 & 1289 & 1260 & 1683 & 1860 \\
&$m_{\tilde{d}_1}$ & 1714 & 2037 & 1043 & 1150 & 1016 & 1316 & 1285 & 1223 & 1678 & 1838 \\
&$m_{\tilde{d}_2}$ & 1651 & 2321 & 1036 & 1341 & 888 & 1379 & 1351 & 1702 & 1688 & 2364 \\
&$m_{\tilde{e}_1}$ & 1532 & 2192 & 970 & 1227 & 769 & 1182 & 1334 & 1696 & 1694 & 2356 \\
&$m_{\tilde{e}_2}$ & 1586 & 2206 & 968 & 1305 & 901 & 1228 & 1256 & 1750 & 1687 & 2370 \\
&$m_{\tilde{\nu}}$ & 1530 & 2196 & 966 & 1296 & 764 & 1202 & 1331 & 1746 & 1692 & 2366 \\
\hline
\multirow{6}{*}{\rotatebox{90}{Other Obs.}}&
$\delta\rho$&\multicolumn{2}{c|}{$8.5\times10^{-6}$}&\multicolumn{2}{c|}{$3.0\times10^{-5}$}&\multicolumn{2}{c|}{$2.3\times10^{-5}$}&\multicolumn{2}{c|}{$2.1\times10^{-5}$}&\multicolumn{2}{c|}{$7.2\times10^{-6}$}\\
&$\delta(g-2)_{\mu}$&\multicolumn{2}{c|}{$6.0\times10^{-11}$}&\multicolumn{2}{c|}{$3.9\times10^{-10}$}&\multicolumn{2}{c|}{$5.5\times10^{-10}$}&\multicolumn{2}{c|}{$7.0\times10^{-11}$}&\multicolumn{2}{c|}{$1.2\times10^{-11}$}\\
&$BR(b\rightarrow s\gamma)$&\multicolumn{2}{c|}{$3.7\times10^{-4}$}&\multicolumn{2}{c|}{$3.9\times10^{-4}$}&\multicolumn{2}{c|}{$3.9\times10^{-4}$}&\multicolumn{2}{c|}{$3.6\times10^{-4}$}&\multicolumn{2}{c|}{$3.7\times10^{-4}$}\\
&$BR(B_s \rightarrow \mu^+\mu^-)$&\multicolumn{2}{c|}{$3.1\times10^{-9}$}&\multicolumn{2}{c|}{$2.7\times10^{-9}$}&\multicolumn{2}{c|}{$2.9\times10^{-9}$}&\multicolumn{2}{c|}{$3.1\times10^{-9}$ }&\multicolumn{2}{c|}{$3.1\times10^{-9}$}\\ \cline{2-12}
&$m_{LMM}$&\multicolumn{2}{c|}{272}&\multicolumn{2}{c|}{175}&\multicolumn{2}{c|}{138}&\multicolumn{2}{c|}{531}&\multicolumn{2}{c|}{487}\\
&$m_{nLMM}$&\multicolumn{2}{c|}{41659}&\multicolumn{2}{c|}{25694}&\multicolumn{2}{c|}{22745}&\multicolumn{2}{c|}{27231}&\multicolumn{2}{c|}{36795}\\
\hline
\hline
\end{tabular}
} }
\end{center}
\caption{\label{tab:weak_obs} \small Weak scale observables, with no
contribution from gauge mediation: $n_3 = n_2 = n_1 = 0$ , see Eqn.
\ref{light_exotica}. We have listed the mass eigenstates of the
squarks and sleptons. Note that for light generations,
$m_{\tilde{u}_1} \approx m_{\tilde{u}_L}$, etc. The last two rows
give the lightest massive modulus ($m_{LMM}$)[mostly K\"{a}hler
modulus (\re{T})] and the \textit{next to} lightest massive modulus
($m_{nLMM}$) [mostly the dilaton (\re{S})]. All other moduli have
mass $\gtrsim 100$ TeV.}
\end{table}
\FloatBarrier
In all cases the lightest MSSM particle is mostly ($\gtrsim 99\%$)
bino (see Table \ref{tab:param_space}). We note that this is
generically true in the models, even when there are contributions
from gauge mediation. The gauge mediated contributions in Eqn.
(\ref{light_exotica}) do not appreciably change the composition of
the LSP, which one can check with the solutions in Table 7 of
Reference \cite{Dundee:2008ts}.
We have evaluated other low energy observables using {\tt
micrOMEGAs} \cite{Belanger:2008sj}. As expected, the bino LSP
overcloses the universe, giving $\Omega_{\textsc{dm}}
>> \Omega_{\textsc{dm}}^{\textsc{obs}}\approx 0.2$. The calculated values
for the following observables are given in the last few rows of
Table \ref{tab:weak_obs}. Corrections to the $\rho$ parameter are
very small. Corrections to $(g-2)_{\mu}$ are significant in Cases 2
and 3 which is not surprising since these are the two cases with the
lightest sleptons for the first two families. We also display the
results for $BR(b\rightarrow s\gamma)$ and $BR(B_s \rightarrow
\mu^+\mu^-)$. The result for $BR(b\rightarrow s\gamma)$ is within
the 2$\sigma$ experimental bound (see \cite{Misiak:2009nr} and
references therein). Given the small chargino masses and the large
values of $\mu$ and the squark and CP odd Higgs masses, we obtain a
branching ratio $BR(B_s \rightarrow \mu^+ \mu^-)$ consistent with
the standard model.
We are not overly concerned about the fact that binos seem to
overclose the universe. In some of the heterotic orbifold models
the Higgs $\mu$ term vanishes in the supersymmetric limit. Hence
there is a Peccei-Quinn symmetry. Supersymmetry breaking effects are
expected to shift the moduli VEVs and generate a non-vanishing $\mu$
term; spontaneously breaking the PQ symmetry and producing the
standard invisible axion. In fact, it has been shown that PQ axions
may be obtained in heterotic orbifold contructions
\cite{Choi:2009jt}. In such cases it is possible that the bino
decays to an axino + photon leaving an axino dark matter candidate
\cite{Rajagopal:1990yx,Covi:1999ty,Covi:2001nw}.
However another, perhaps more important, cosmological effect must be
considered. All 5 cases have a gravitino with mass less than 3
TeV. Thus there is most likely a gravitino problem. In addition
the lightest moduli mass is of order (Table \ref{tab:weak_obs})
several 100s GeV. Thus there is also a cosmological moduli problem.
But there is hope. The next lightest massive modulus [nLMM] has, in
all cases, a mass above 20 TeV. A detailed cosmological analysis
is beyond the scope of this paper. However, it is possible that
when cosmological temperatures are of order $m_{\text{nLMM}}$, the
universe becomes nLMM dominated. By the time the nLMM decays all
matter is diluted and then the universe reheats to temperatures
above the scale of big bang nucleosynthesis (for example, see
\cite{Acharya:2008bk}). Thus it is possible that the nLMM solves
both the gravitino and light moduli problems. Of course, then the
issue of obtaining the correct baryon asymmetry of the universe and
the dark matter abundance must be addressed. Both can in principle
be obtained via non-thermal processes at low temperature.
In Table \ref{tab:param_space} we analyze the dependence of our
results on the value of $\tan\beta$ and $\text{sgn}(\mu)$ with all
other input parameters fixed. We find that only the value of the
light Higgs mass is sensitive to varying $\tan\beta$. Note the
lowest value of $\tan\beta$ is obtained by the Higgs mass bound,
while the largest value of the light Higgs mass is obtained with the
largest value of $\tan\beta$ (for both signs of $\mu$).
Additionally, at large $\tan\beta$ for $\mu < 0$ the Higgs potential
becomes unbounded from below. For $\mu > 0$ we limited the analysis
to $\tan\beta \leq 50$. The light Higgs mass does not go above 122
GeV for $\tan\beta \leq 50$.
\begin{table}[t!]
\renewcommand\arraystretch{1.2}
\begin{center}
\noindent\makebox[\textwidth]{ \scalebox{0.75}{
\begin{tabular}{|cc|c|c|c|c|c|c|c|c|c|c|}
\hline
\multicolumn{2}{|c|}{\multirow{2}{*}{Case}}&\multicolumn{2}{c|}{1}&\multicolumn{2}{c|}{2}&\multicolumn{2}{c|}{3}&\multicolumn{2}{c|}{4}&\multicolumn{2}{c|}{5} \\
\cline{3-12}
&&$\mu < 0$&$\mu > 0$&$\mu < 0$&$\mu > 0$&$\mu < 0$&$\mu > 0$&$\mu < 0$&$\mu > 0$&$\mu < 0$&$\mu > 0$ \\
\hline
\multirow{2}{*}{$\tan\beta$} & lo & 5 & 6 & 8 & 12 & 9 & 11 & 5 & 4 & 5 & 3 \\
& hi & 38 & 50 & 36 & 50 & 39 & 50 & 32 & 48 & 39 & 50 \\
\hline
\multirow{2}{*}{$m_{h^0}$} & lo & 113.5 & 113.2 & 112.4 & 112.4 & 112.4 & 112.4 & 112.4 & 116.2 & 113.5 & 112.5 \\
& hi & 117.4 & 117.2 & 113.7 & 113.4 & 113.6 & 113.7 & 120.5 & 121.8 & 120.8 & 121.9 \\
\hline
Neut. & & bino & bino & bino & bino & bino & bino & bino & bino & bino & bino \\
comp. & & $\gtrsim 99\%$ & $\gtrsim 99\%$ & $\gtrsim 99\%$ & $\gtrsim 99\%$ & $\gtrsim 99\%$ & $\gtrsim 99\%$ & $\gtrsim 99\%$ & $\gtrsim 99\%$ & $\gtrsim 99\%$ & $\gtrsim 99\%$ \\
\hline
$m_{\tilde{\chi}^0_1}$ & lo & 149.1 & 148.5 & 82.5 & 82.0 & 98.6 & 98.2 & 68.9 & 67.3 & 53.6 & 51.5 \\
(GeV) & hi & 151.5 & 149.9 & 84.0 & 82.9 & 99.8 & 98.7 & 69.6 & 70.3 & 55.1 & 55.9 \\
\hline
$m_{\tilde{\chi}^{\pm}_1}$ & lo & 290.4 & 291.3 & 162.3 & 162.3 & 193.5 & 193.6 & 139.4 & 134.3 & 110.3 & 103.3 \\
(GeV) & hi & 298.8 & 293.7 & 167.1 & 163.7 & 197.7 & 194.5 & 141.0 & 141.6 & 113.8 & 114.6 \\
\hline
\hline
\end{tabular}
}} \caption{\label{tab:param_space}Scan over $\tan\beta$ and
$\text{sgn}(\mu)$.}
\end{center}
\end{table}
\FloatBarrier
\section{Conclusions}
\label{sec:conclusion}
As a candidate theory of all fundamental interactions, string theory
should admit at least one example of a four-dimensional vacuum which
contains particle physics and early universe cosmology consistent
with the two standard models. In this context, the recently found
``mini-landscape" of heterotic orbifold
constructions~\cite{Buchmuller:2005jr, Buchmuller:2006ik,
Lebedev:2006kn, Lebedev:2007hv, Lebedev:2008un} provide us with very
promising four-dimensional perturbative heterotic string vacua.
Their low-energy effective field theory was shown to resemble that
of the MSSM, assuming non-zero VEVs for certain blow-up moduli
fields which parametrize resolutions of the orbifold fixed points
along $F$- and $D$-flat directions in global supersymmetry.
In this paper we have dealt with the task of embedding the globally
supersymmetric constructions of the heterotic ``mini-landscape" into
supergravity and then stabilizing the moduli of these
compactifications, including their orbifold fixed point blow-up
moduli. The blow-up moduli appear as chiral superfields contained in
the twisted sectors of the orbifolded heterotic string theory. They
are singlets under all standard model gauge groups, but are charged
under several unwanted $U(1)$ gauge symmetries, including the
universal anomalous $U(1)_A$ gauge symmetry of the heterotic string.
Note, moduli stabilization of string compactifications is a crucial
precondition for comparing to low energy data, as well as for
analyzing any early universe cosmology, such as inflation, in a
given construction.
Section \ref{sec:general_structure} served the purpose of reviewing
the ingredients and structure of the heterotic 4d $N=1$
supergravity inherited from orbifold compactifications of the 10d
perturbative $E_8\otimes E_8$ heterotic string theory. The general
structure of these compactifications results in:
\begin{itemize}
\item[i)] a standard no-scale K\"{a}hler potential for the bulk volume and
complex structure moduli, as well as the dilaton, together with
\item[ii)] gaugino condensation in the unbroken sub-group of the hidden $E_8$, and
\item[iii)] the fact that the non-perturbative (in the world-sheet instanton sense)
Yukawa couplings among the twisted sector singlet fields contain
terms explicitly breaking the low-energy $U(1)_R$-symmetry.
\end{itemize}
We have shown in Section \ref{sec:one_gaugino_condensate} that these
three general ingredients, present in all of the ``mini-landscape"
constructions, effectively realize a KKLT-like setup for moduli
stabilization. Here, the existence of terms explicitly breaking the
low-energy $U(1)_R$-symmetry at high order in the twisted sector
singlet fields is the source of the effective small term $w_0$ in
the superpotential, which behaves like a constant with respect to
the heterotic dilaton \cite{Kappl:2008ie}. Utilizing this, the
presence of just a single condensing gauge group in the hidden
sector (in contrast to the racetrack setups in the heterotic
literature) suffices to stabilize the bulk volume $T$ (and, by
extension, also the bulk complex structure moduli $U$), as well as
the dilaton $S$ at values $\langle {\rm Re}\;T\rangle \sim 1.1 -
1.6$ and $\langle{\rm Re}\;S\rangle\sim 2$. These are the values
suitable for perturbative gauge coupling unification into $SU(5)$-
and $SO(10)$-type GUTs distributed among the orbifold fixed points.
Note, we have shown this explicitly for the case one $T$ modulus and
a dilaton, however, we believe that all bulk moduli will be
stabilized near their self-dual points
\cite{Font:1990nt,Cvetic:1991qm}.
At the same time, the near-cancelation of the $D$-term of the
universal anomalous $U(1)_A$-symmetry stabilizes non-zero VEVs for
certain gauge invariant combinations of twisted sector singlet
fields charged under the $U(1)_A$. This feature in turn drives
non-vanishing $F$-terms for some of the twisted sector singlet
fields. Thus, together with the $F$-terms of the bulk volume moduli
inherited from modular invariance, it is sufficient to uplift the
AdS vacuum to near-vanishing cosmological constant.
The structure of the superpotential discussed in this paper, $\ensuremath{\mathcal{W}}
\sim w_0 e^{- b T} + \phi_2 \ e^{- a S - b_2T}$, behaves like a
`hybrid KKLT' with a single-condensate for the dilaton S, but as a
racetrack for the $T$ and, by extension, also for $U$ moduli. An
additional matter $F_{\phi_2}$ term driven by the cancelation of the
anomalous $U(1)_A$ $D$-term seeds successful up-lifting.
We note the fact that the effective constant term in the
superpotential, $w_0$, does not arise from a flux superpotential akin to the
type IIB case. This leaves open (for the time being) the question
of how to eventually fine-tune the vacuum energy to the
$10^{-120}$-cancelation necessary.
Section \ref{sec:moduli_stabilization} then serves to demonstrate
how the success of stabilizing the bulk moduli and breaking
supersymmetry in the $F$-term sector, driven by the $U(1)_A$ $D$-term
cancelation, transmits itself to the chiral singlet fields from the
untwisted and twisted sectors of the orbifold compactification which
contain, among others, the blow-up moduli associated with the
orbifold fixed points. The effects from the bulk moduli
stabilization and supersymmetry breaking, transmitted through
supergravity, generically suffice to stabilize all of the twisted
sector singlet fields at non-zero VEVs. This property was assumed
in the original ``mini-landscape" construction in order to decouple
the non-MSSM vector-like exotic matter, and our arguments provide the
first step towards a self-consistent justification for these
assumptions.
In Section \ref{sec:spectrum} we estimate the structure of the soft
terms from the moduli sector supersymmetry breaking at the high
scale. We find that the contributions from high-scale gauge
mediation are subdominant (although not parametrically suppressed)
compared to the gravity mediated contributions. Upon RGE running the
high-scale soft terms to the weak scale using {\tt softSUSY}, we
obtain several benchmark patterns of sparticle and Higgs masses (see
Table \ref{tab:weak_obs}). The low-energy spectrum features an
allowed window of $\tan\beta$ values for $m_{3/2}<5\,{\rm TeV}$. It
generically contains a light chargino/ neutralino spectrum and heavy
squarks and sleptons. The lightest MSSM partner, in the 5 benchmark
cases studied, is given by a bino ($> 99\%$) with mass $\gtrsim 52
\,{\rm GeV}$. If this were the LSP, it would yield a dark matter
abundance which over closes the universe, however, the
``mini-landscape'' models offer some possible resolutions. One
possibility is that the bino decays into an axino, the partner of
the invisible axion responsible for canceling the $\theta$-angle of
QCD, which is present in many of the ``mini-landscape" setups
\cite{Choi:2009jt}. We have also considered an alternative
possibility that the late decay of the next to lightest massive
modulus might ameliorate or solve the cosmological gravitino and
moduli problem. This would then dilute the above mentioned
cosmological abundance of binos. Of course, the non-thermal
production of dark matter and a baryon asymmetry must then be
addressed. Note, however, the resolution of these cosmological
questions are beyond the scope of the present paper.
Summarizing, we have given a mechanism for moduli stabilization and
supersymmetry breaking for the perturbative heterotic orbifold
compactifications. It relies on the same variety and number of
effective ingredients as the KKLT construction of type IIB flux
vacua and thus represents a significant reduction in necessary
complexity, compared to the multi-condensate racetrack setups
utilized so far. When applied to a simplified analog of the
``mini-landscape" heterotic orbifold compactifications, which give
the MSSM at low energies, it leads to fully stabilized 4d heterotic
vacua with broken supersymmetry and a small positive cosmological
constant. Moreover, most of the low energy spectrum could be
visible at the LHC.
We leave some important questions like the problem of the full
fine-tuning of the vacuum energy to near-vanishing, or the existence
of an inflationary cosmology within these stabilized
``mini-landscape" constructions for future work. Further study is also warranted with respect to potential cosmological moduli and gravitino problems that may be associated with sub-100 TeV moduli and gravitino mass values (see e.g.~\cite{gravitinoprob}). Finally, the
numerical evaluation of any particular ``mini-landscape" vacuum
requires analyzing the supergravity limit with three bulk moduli,
$T$, one bulk complex structure modulus, $U$, and of order 50
blow-up moduli. A detailed analysis of this more realistic situation
would require a much better handle on the moduli space of heterotic
orbifold models than is presently available.
\newpage
\section*{Acknowledgements}
A.W. would like to thank the Physics Department at the Ohio State
University for their warm hospitality, where part of this work was
completed. B.D. and S.R. would like to thank the Stanford Institute
for Theoretical Physics for their hospitality, where the bulk of
this work was completed. Also B.D., S.R. and A.W. thank the KITP
at UC Santa Barbara for support. We are also indebted to Patrick
Vaudrevange for aid with the superpotential for Model IA. B.D. and
S.R. are supported by DOE grant DOE/ER/01545-884. A.W. is supported
in part by the Alexander-von-Humboldt foundation, as well as by NSF
grant PHY-0244728. This research was supported in part by the National Science Foundation under Grant No.~NSF PHY05-51164. We would also like to thank K.~Bobkov and
N.~Craig for useful discussions. S.R. would also like to thank
W.~Buchm\"{u}ller, J.~Conlon, E.~Dudas, G.~Dvali, T.~Kobayashi, D.~L\"{u}st, H.P.~Nilles, M.~Ratz, S.~Ramos-Sanchez, and G.G.~Ross for discussions.
\clearpage\newpage
|
\section{Introduction} \label{intro}
The question of what is the central density profile of galactic dark
matter haloes has been much debated in the literature over
many years. Since the work of Navarro et. al (1997), cosmological
N-body simulations have consistently agreed in yielding central
density profiles which can be accurately fitted by functional forms
characterised by centrally divergent density cusps. Although the
details vary and the innermost slope and radius to which
said fits are to be trusted are still discussed, a broad agreement has
been reached in that cosmologically simulated dark haloes exhibit what
are termed 'cuspy density profiles', e.g. Merritt et al. (2006).
On the other hand, observational inferences of dark halo density
profiles through rotation curve decomposition, have tended to favour
those showing density profiles which tend to constant
values towards the centre. Recent examples include de Blok et
al. (2008), Kuzio de Naray et al. (2009) and Gebhardt \& Thomas
(2009). The issue is complicated by the necessity to estimate the
dynamical relevance of the baryonic component, a function of the
assumed mass to light ratio, the relevance of observational
uncertainties such as beam smearing, and the
importance of non circular motions and non centrifugal support of
asymmetric drift and hydrodynamic pressure terms (e.g. Valenzuela et
al. 2007).
An interesting independent clue to the puzzle might come from the
presence of massive black holes in the centre of dark matter
haloes. The relevance of such black holes is well established from
their role as the central engines for quasars and active galaxies
(Rees 1984) as well as in quiescent systems (Kormendy \& Richstone
1995). Although mostly seen as active nuclei in the distant universe,
it is commonly believed that all large galaxies host such objects
in their centres. Recent empirical determinations of central QSO
black hole masses have established their existence with masses in the
range $10^{7} - 10^{10} M_{\odot}$ at redshifts beyond $z\simeq 3$
(Kelly et al. 2008, Graham 2008).
We therefore must conclude that the central regions of large dark
haloes have coexisted with massive black holes over most of the
history of the universe. Given the existence of event horizons
associated to black holes, and the assumption of standard cold dark matter
subject only to gravitational interactions, it follows that central
black holes have grown over the history of galactic dark haloes,
through the capture of dark matter particles.
The problem of the growth of single central galactic black holes
through accretion was addressed by Lynden-Bell \& Rees
(1971) and more recently by Gnedin \& Primack (2004) and Zhao et
al. (2002), but only in the accretion of particles on capture orbits.
While readily acreted, they constitute only a
minor fraction of those available in the distribution function of halo
particles. Further, once absorbed, one has to wait over a
comparatively long halo relaxation timescale for it to be re-populated
with dark matter particles. Here we consider only the accretion of
unbound particles, through the absorption
cross section presented by the black hole through its event horizon,
enhanced by gravitational focusing. In this case, the accretion is
slower at first, but proceeds at an ever increasing rate as the
growing black hole mass leads to an increase in its area, with little
accompanying depletion of the overall dark matter halo distribution
function. This, as accretion does not takes place over a highly
specific fraction of the distribution function, while the black hole
mass constitutes only a fraction of the overall dark halo mass.
We find the process to be
characterised by the onset of a rapid runaway growth phase after
a critical timescale. This timescale is a function of
the mass of the black hole and the local density of dark matter.
By requiring that the runaway phase does not occur, as then the
swallowing up of the halo by the black hole would seriously distort
the former, we can obtain upper limits to the maximum allowed
density of dark matter at the centres of haloes.
\section{Central Black Hole Growth Rates}\label{analytical}
In the case of cuspy dark matter profiles, the scales over which the
density varies by a significant factor, even in the central regions,
are typically of order $10$~pc or above (Merritt et al. 2006,
Stadel et al. 2009). This is many orders of magnitude greater than the
typical length scales over which the accretion onto the central black
hole of mass $M$ takes place, the Schwarzschild radius, $R_{\rm Sch}=5
\times 10^{-4}{\rm pc} (M/5 \times 10^{9} M_{\odot})$. The range of
scales involved, from $R_{\rm Sch}$ to the scale of the dark matter
halo, make a full simulation of the whole problem unfeasible. To
first approximation, we will therefore treat the central region of the dark matter
halo over which the black hole finds itself as one of constant density.
The effects of central black holes, mostly binary ones, on the stellar
population of a galaxy have been treated before, e.g Quinlan (1996),
Merritt \& Milosavljevic (2005), Sesana et al. (2007). The problem
here is different because the total mass of the dark halo, and the
range of radii covered by the dark matter particles are both much
larger than equivalent quantities in the case of bulge stars affected
by central black holes. Since the mass of the black hole is still
several orders of magnitude smaller than the total mass of the dark
halo, we shall treat the presence of the black hole as a perturbation
on the distribution function of the dark matter particles.
In Hernandez \& Lee (2008), we calculated the density response of a
constant density, isothermal dark matter distribution, to the presence
of a point mass $M$. A highly localised cusp appears, with the density
profile changing from $\rho_{0}$ to $\rho_{0}+\rho_{1}(r)$, where
\begin{equation}
\rho_{1}(r)=\frac{G M}{r \sigma^{2}} \rho_{0}.
\end{equation}
In the above, $\sigma$ is the velocity dispersion of the halo
particles, and $r$ is the distance to the mass $M$. In estimating the
growth rate of the central black hole, of initial mass $M_{0}$, we
shall begin by assuming that its presence will elicit a response in
the otherwise unperturbed dark halo particles, as described by
eq.~(1). In Hernandez \& Lee (2008) we showed
through direct comparison with high resolution N-body simulations,
that the analytic expression in eq.~(1) accurately
describes the response of a constant density isothermal region to the
presence of a point mass. A first correction due to relativistic
effects, the substitution of a Newtonian potential for the expression
of Paczy\'{n}ski \& Wiita (1980)\footnote{In the P-W expression a point
mass produces a gravitational field $\Phi(r)=-GM/(r-R_{\rm Sch})$
instead of the usual $\Phi=-GM/r$.}, will result only in the
substitution of $r-R_{\rm Sch}$ for the current $r$ in the denominator of
eq. (1).
We can now estimate the growth rate of the central black hole
dimensionally as:
\begin{equation}
\dot{M}=C_{0} \rho A \sigma,
\end{equation}
where $\rho$, $A$ and $\sigma$ are a characteristic density, area and
velocity for the spherical accretion in question, and $C_{0}$ is a
dimensionless constant which one would expect to be of order
unity. From the preceding discussion regarding the reaction of the
halo to the presence of the black hole, and from the fact that
accretion of unbound particles will occur on crossing the event
horizon at $R=R_{\rm Sch}$, we can estimate $\dot{M}$ from taking
$\rho = \rho_{1}(R_{\rm Sch})$, $A=4 \pi R_{\rm Sch}^{2}$ and $\sigma$
as the velocity dispersion of the dark matter particles in the
halo. We have ignored $\rho_{0}$ in favour of $\rho_{1}$ in the above
considerations, as at distances of order $R_{Sch}$ the latter dominates
over the former by a factor of order $(c/\sigma)^{2}$. The above yields:
\begin{equation}
\dot{M}=C_{0} 8 \pi \frac{G^{2} M^{2}\rho_{0}}{\sigma c^{2}}.
\end{equation}
Gravitational focusing effectively increases the cross section of the
black hole by trapping particles which would otherwise fail to be
acreted, and enters into the computation of $C_{0}. $In fact, a fully
relativistic calculation for the accretion rate of a black hole
immersed in an isothermal distribution of non-relativistic particles
leads to the result:
\begin{equation}
\dot{M}=16 \left( 6 \pi \right)^{1/2} \frac{G^{2} M^{2}\rho_{0}}{\sigma c^{2}},
\end{equation}
as derived in Shapiro \& Teukolsky (1983), eq.~14.2.26, for particles with positive energies.
Those on bound orbits can be engulfed rapidly in a short initial transient phase, and will
re-appear as the corresponding phase space is re-populated by the
distribution function on the very long relaxation timescales of the
full halo (Shapiro \& Teukolsky 1983). Here again,
by considering only the accretion rate of unbound particles, we are
confident in having a secure lower limit on the accretion
rate. Comparing with the dimensional analysis of eq.~(3) we see
that the result is exact for $C_{0}=2(6/ \pi)^{1/2}$, a factor of less
than 3. In what follows we shall use the exact result of eq.~(4),
with eq.~(3) serving only in allowing a physical
interpretation of the relativistic result.
Introducing dimensionless quantities $\Sigma=\sigma/c$, $\tau=t/t_{\rm ff}$
and ${\mathcal M}=M/M_{\rm J}$, where $t_{\rm ff}= 1/(G \rho_{o})^{1/2}$ is the free
fall timescale of the unperturbed background density,
and $M_{\rm J}=\sigma^{3}/(G^{3/2} \rho_{0}^{1/2})$
is the Jeans mass of the unperturbed halo, we obtain the dimensionless
growth rate:
\begin{equation}
\frac{d{\mathcal M}}{d \tau} = (6 \pi)^{1/2}(4 {\mathcal M} \Sigma)^{2}.
\end{equation}
We note that in cosmological N-body simulations, the distribution
function of dark matter particles exhibits a large
degree of orbital anisotropy and is dominated by highly radial orbits
(e.g. Ascasibar \& Gottlober 2008). The increased fraction of the dark
halo hence available for interaction with the central black hole, and
the reduced angular momentum of the dark matter particles, compared to
the isothermal case, will all tend to yield faster growth
rates than those calculated here. The upper limits on central
density derived below are hence safe upper estimates. More detailed
calculations accounting for an intrinsically cusped dark halo profile
with radially dominated distribution functions would yield even lower
limit densities.
\section{Central Dark Matter Density Limits}
From eq.~(4) we obtain for the black hole mass:
\begin{equation}
M(t)=\frac{M_{0} c^{2} \sigma}{c^{2}\sigma - 16(6 \pi)^{1/2} G^{2}\rho_{0} M_{0} t}.
\end{equation}
There is a strong divergence for $t \rightarrow
t_{\rm div} = c^{2} \sigma /(16(6 \pi)^{1/2} G^{2} \rho_{0} M_{0})$. The time
for it to appear decreases as the
initial black hole mass rises, as the central dark matter density
increases, and increases as the velocity dispersion of the halo
particles rises. The divergence is so abrupt that the
time it takes for the black hole mass to increase by one order of
magnitude, $T_{10}$, is only 9/10 $t_{\rm div}$. The evolution we
calculate will still be accurate up to $t=T_{10}$, as the total mass
of the dark halo (several times $10^{12} M_{\odot}$ for large galactic
haloes), will still be over an order of magnitude larger than that of
the central black hole, even for the largest initial black hole masses
considered, of a few times $10^{9} M_{\odot}$. We shall therefore define:
\begin{equation}
T_{10} = \left( \frac{9}{10} \right) \frac{c^{2} \sigma}{16 (6 \pi)^{1/2} G^{2} \rho_{0} M_{0}},
\end{equation}
as a characteristic timescale after which the accretion process
results in substantial dynamical alterations to the overall dark
halo. With the same dimensionless quantities as defined previously,
we obtain the corresponding expressions:
\begin{equation}
{\mathcal M}(\tau)=\frac{{\mathcal M}_{0}}{1-C_{1}{\mathcal M}_{0} \Sigma^{2} \tau},
~~~~~~~ \tau_{\rm div}=\left(C_{1} {\mathcal M}_{0} \Sigma^{2} \right)^{-1},
\end{equation}
where $C_{1}=16(6 \pi)^{1/2}$.
We can now calculate the evolution of eq. (6) for any
value of the central black hole mass. We begin with parameters as
appropriate for the largest inferred QSO central black holes, $M_{0}=5
\times 10^{9} M_{\odot}$ (Kelly et al. 2008, Graham 2008), as this
case will lead to the most restrictive dark matter density
limits. Although the dark haloes of QSOs cannot be observationally
inferred, given the scalings observed at low redshift between black
hole masses and galactic properties (e.g. Kormendy \& Richstone 1995,
Gebhardt et al. 2000, Ferrarese \& Merritt 2000, Tremaine et al. 2002,
Gultekin et al. 2009), between black hole masses and halo properties
(e.g. Volonteri et al. 2003, Bolton et al. 2008, Croton 2009, Bandara
et al. 2009) and between galactic
properties and dark halo masses, such as the Tully-Fisher relation
(Tully \& Fisher 1977), it is reasonable to assume that QSOs hosting
the largest inferred black hole masses will be hosted by large
galactic haloes. Consequently, we take a large value of $\sigma=200$
~km/s. Given the usual scaling between $\sigma$ and the flat rotation
curve velocity of a galactic halo of $V_{rot} = 2^{1/2} \sigma$, this choice corresponds to $V_{rot} = 280$~km/s, a
value in the extreme range for any type of galactic system. Notice
that as $t_{\rm div}$ scales with $\sigma$, taking a large value for
this parameter will again result in conservative upper limits on the
final inferred limit central halo densities (see below). Also, given
the 'inside out' and 'downsizing' aspects of current cosmological
structure formation models, (e.g. Naab et al. 2009) the dynamical
stability of the central regions of the most massive galactic systems
over the lifetimes we have assumed appears reasonable.
Figure~1 shows the growth of a central black hole as a function
of time, for our fiducial case with $M_{0} =5 \times 10^{9} M_{\odot}$
and $\sigma=200$~km/s, and a range of values for the assumed central
dark matter density $\rho_{0}$=400, 350, 300, 250, 200, 150 and 100
$M_{\odot}$~pc$^{-3}$ (from top to bottom, respectively). The
case where $T_{10}=10$~Gyr corresponds to the middle curve, where the
central dark matter density is $\rho_{0}=250 M_{\odot}$~pc$^{-3}$. The
mass of the black hole increases by a factor of 10 in 10~Gyr, a
conservative estimate of the lifetime of the systems in question,
namely, dark haloes of QSOs observed at high redshift.
We see that for central dark matter densities above this threshold of
$250 M_{\odot}$~pc$^{-3}$, the mass of the central black hole enters
the runaway accretion regime and diverges on timescales shorter than
the lifetimes of the systems being treated, as given by the three
upper curves in Figure~1. On the
other hand, for values below this threshold,
the growth of the central black hole is of only a factor of order
unity over 10~Gyr, as shown by the three lower curves.
\begin{figure}
\includegraphics[angle=0,scale=0.4]{fig1.ps}
\caption{Evolution of the mass of a central black hole with initial
mass $M_{0}=5 \times 10^{9} M_{\odot}$ in a dark halo with dark matter
particles of isotropic velocity dispersion $200$~km/s, for
varying central region dark matter density:
$\rho_{0}$=400, 350, 300, 250, 200, 150 and 100
$M_{\odot}$~pc$^{-3}$, top to bottom, respectively.}
\label{fig:picture}
\end{figure}
For this particular initial mass, $M_0=5 \times 10^{9}M_{\odot}$, we
can hence identify $250 M_{\odot}$~pc$^{-3}$ as a maximum central halo
dark matter density above which the inferences of black hole masses in
high redshift QSOs would imply growth rates for the central black
holes resulting in substantial dynamical distortions, leading today
not to quiescent black holes in the centres of normal galaxies, but to
exotic objects dynamically dominated by extreme super massive black
holes. Consistency arguments of this type can be found e.g. in Gnedin
\& Ostriker (2001), who calibrate the physical parameters of self-interacting
dark matter by requiring that galactic dark haloes should not have evaporated
by now into galaxy cluster dark matter haloes. If we were to take larger
values for the black hole mass, such as those given by Graham (2008),
reaching $10^{10} M_{\odot}$, or upwards of $10^{10} M_{\odot}$ reported
for some objects by Kelly et al. (2008), the threshold density we
identify would go down by a factor of a few.
In Figure~2 we show the constraints on the central dark halo
densities, $\rho_{M}$, as a function of the assumed central black hole
mass and for a fixed value of the velocity dispersion,
$\sigma=200$~km/s. The curves correspond to various values of $T_{10}$.
By taking a conservative measure of the lifetimes of the systems in question as
10~Gyr, the region containing the dotted curves above the thick black
line at $T_{10}=10$~Gyr is excluded from consistency arguments, while
the allowed region of parameter space lies in the half plane below it.
The choice of values for the lifetimes of high redshift
QSOs larger than the 10~Gyr previously assumed, shifts the maximum
central dark matter density values downwards onto the various thin
solid curves. The most stringent limits apply to the
highest black hole masses at $M \geq 5 \times
10^{9}M_{\odot}$. However, the expectation of universality for the
cosmological dark matter density profiles leads one to expect these
limits will apply to all dark matter haloes.
\begin{figure}
\includegraphics[angle=0,scale=0.4]{limits.ps}
\caption{Limit central densities as a function of central black hole
masses, for a number of values of $T_{10}$ going from 1 to 15 Gyr,
every Gyr, from top to bottom respectively. The region above the
thick black line at $T_{10}=10$~Gyr, with dotted lines, is excluded
from consistency arguments.}
\label{fig:picture}
\end{figure}
We note that once the mass of the central black hole grows
substantially, processes not included here would begin to
become relevant, and would invalidate the simple physical
hypothesis leading to eq.~(4). Some include: the
adiabatic contraction of the dark halo in response to the
concentration of mass into the central black hole, resulting in
higher central dark matter densities and hence even higher accretion
rates; the accretion of a fraction of the baryons into the central
black hole, which is known to occur; or the enhanced
gravitational focusing of matter of all types into the black hole,
once the approximation of the black hole mass being small compared to
the total halo mass which we are working under begins to break
down. All of these make it reasonable to assume that
the first corrections to eq.~(4) will lead to even larger
accretion rates, hence leaving our conclusions, in terms of limit
densities, unchanged.
Regarding the accretion of baryons and dark halo particles, it has
been proposed that this can be partly responsible for the
appearance of the observed scaling laws between central black hole
masses and bulge and galactic properties, e.g., in the analytical work
of Zhao et al. (2002) and Gnedin \& Primack (2004), the large scale
simulations of Di Matteo et al. (2008), or the luminosity function
consistency arguments in Yu \& Tremaine (2002) and Hopkins et
al. (2007). Indeed, Hennawi \& Ostriker (2002) constrain the velocity
dependence of the interaction cross-section of hypothetical
self-interacting dark matter, by requiring that the accretion of dark
matter onto central black holes leads to the observed $M-\sigma$
relation. Also for the case of self-interacting dark matter, Balberg
\& Shapiro (2002) explore the formation of supermassive black
holes through gravothermal core-collapse of the central regions of
galactic dark haloes.
Comparing the upper limiting central dark matter density of
$250 M_{\odot}$~pc$^{-3}$ with the dynamically inferred structure of galactic dark
haloes, it is reassuring that when a constant density core is
used to model observations, the inferred central dark matter
densities always lie below this limit, typically at $\simeq 1
M_{\odot}$~pc$^{-3}$, or below. Recent examples are given by
Gilmore et al. (2007) for local dwarf spheroidal
galaxies, and de Blok et al. (2008) for late type galaxies.
Hence no conflict appears, in that the runaway accretion regime for
the central black hole will not be reached in 10 Gyr for any
directly inferred values of the central dark matter density, for any
inferred central black hole masses.
From the point of view of cuspy dark matter haloes, the limits we
derive here establish an inner boundary, exterior to which the
globally fitted centrally divergent dark halo profiles can be
valid. At smaller radii, this solution must be modified to avoid the
divergent black hole growth rates found here. Although our high limit
densities are not reached by current cosmological N-body simulations,
which typically stop at volume densities of order $1
M_{\odot}$~pc$^{-3}$ at their resolution limit, even for the most
recent highest resolution experiments (Cuesta et al. 2008, Stadel et
al. 2009), the logarithmic slopes in these regions are still such that
volume densities above the limits we derive here would be reached
orders of magnitude in the radial coordinate before reaching the scale
of the super massive central black holes, even for central black hole
masses below the upper ranges of these values. Studies of
the origin of cuspy cosmological density profiles
within the secondary infall scenario have traced the cusp to the close
to scale free initial perturbation spectrum (e.g. Williams et al. 2004,
Salvador-Sol\'{e} et al. 2005, Del Popolo 2009). Such studies explain the steep negative
logarithmic slopes of density profiles from N-body simulations,
predicting them to extend into the very centres. Although these
considerations apply only to cosmological dark matter haloes in the
absence of any central black holes, the inclusion of single central
black holes, currently beyond the reach of fully self consistent
simulations, will result in even steeper central dark matter profiles,
strengthening the consistency arguments made here. A solution
might include dark matter physics not ordinarily
considered, such as self-interacting dark matter
(e.g. Firmani et al. 2000, de la Macorra 2009), warm dark
matter, or changes to the initial fluctuation spectrum
(e.g. Alam et al. 2002).
\section{Conclusions}\label{ccl}
We study the mass growth rates of central black holes through
accretion of dark matter particles on orbits unbound to the central
black hole. As the black hole mass grows, a runaway
accretion regime ensues. Requiring that no such runaway regime has
been reached over lifetimes of galactic dark haloes of 10~Gyr leads
to the identification of critical upper limits for
the density of dark matter. These limits
scale proportionally to the assumed value of the dark matter velocity dispersion, and
inversely proportionally to the assumed value of the central black hole mass.
For the largest black hole masses inferred by QSO studies of
$5 \times 10^{9} M_{\odot}$, central region dark matter densities larger than
$\rho_{0} = 250 M_{\odot}$~pc$^{-3}$ are excluded.
These limits suggest dark halo density structures are characterised by
constant density central regions, rather than divergent cuspy profiles.
\section*{acknowledgements}
The authors acknowledge the constructive criticism of an anonymous
referee in helping to reach a more clear and complete final version.
Financial support for this work was provided in part through CONACyT
(83254E), and DGAPA-UNAM (PAPIIT IN-113007-3 and IN-114107).
|
\section{Introduction}
\label{sec:introduction}
Choreographies are global descriptions of transactions including
business or financial transactions. They describe the intertwined
behavior of several principals as they negotiate some agreement
and--frequently--commit some state change. A key idea is
\emph{end-point projection}~\cite{carbone.honda.yoshida:esop07}, which
converts a global description into a set of descriptions that
determine the local behavior of the individual participants in a
choreography. Conversely, \emph{global synthesis} of a choreography
from local behaviors is also sometimes possible, i.e.~meshing a set of
local behaviors into a comprehensive global
description~\cite{MostrousYoshidaHonda09}.
Because these transactions may transfer sums of money and other
objects of value, or may communicate sensitive information among the
principals, they require a security infrastructure. It would be
desirable to synthesize a cryptographic protocol directly from a
choreography description, to control how adversaries can interfere
with transactions among compliant principals. Corin et
al.~\cite{CDFBL08} have made a substantial start on this problem, with
further advances described in~\cite{BhargavanEtAl09}. However, many
questions remain, for instance how to optimize the generated
cryptographic protocols, how best to establish that they are always
correct, and indeed how best to define their correctness.
This last question concerns how to state what control the protocol
should provide, against adversaries trying to interfere with
transactions. It is a substantial question because the execution
model that choreographies use is quite distant from the execution
model cryptographic protocols are designed to cope with. For
instance, choreographies use an execution model---or communication
model---in which messages are never received by any party other than
the intended recipient, or if the formalism represents channels, they
are received only over the channel. Moreover, messages are always
delivered if the recipient is willing to receive the message.
Messages are delivered only if they were sent, and specifically only
if they were sent by the expected peer. Finally, they are delivered
only once. These aspects of the model mean that confidentiality and
integrity properties are built into the underlying assumptions. A
security infrastructure is intended to justify exactly these
assumptions, i.e.~to provide a set of behaviors in which these
assumptions are satisfied.
Naturally, these behaviors must be achieved within an underlying model
in which the adversary is much stronger. In this model---typically
called the \emph{Dolev-Yao model}, after a paper~\cite{DolevYao83} in
which Dolev and Yao formalized ideas suggested by Needham and
Schroeder~\cite{NeedhamSchroeder78}---all messages may be received by
the adversary, so that confidentiality needs to be achieved by
encryption. They may be delivered zero times, once, or repeatedly,
and they may be misdelivered to the wrong participant. When
delivered, a message may appear to come from a participant that did
not send it. The adversary may alter messages in transit, including
applying cryptographic operations using keys that he knows, or may
obtain by manipulating the protocol.
Digital signatures may be used to notify a recipient reliably of the
source of a message (and of the integrity of its contents). Symmetric
encryption may also be used to ensure authenticity: a recipient knows
that the encrypted message was prepared by a party that knew the
secret key, and intended it for a peer that also knew the secret key.
Nonces, which are simply randomly chosen bitstrings, may be used to
ensure freshness. The principal $P$ that chose a nonce knows, when
receiving a message containing it, that the nonce was inserted after
$P$ chose it. Moreover, if $P$ engages in many sessions and
associates a different nonce with each, $P$ can ensure that messages
containing one nonce cannot be misdirected to a session using a
different nonce.
In this paper, we begin the process of relating the Dolev-Yao model of
execution to the choreography execution model. This is a key step in
generating cryptographic protocols and proving them faithful to the
intent of the choreography. In particular, we represent the two
execution models using the strand space
model~\cite{strandspaces,GuttmanEtAl05}.
\paragraph{Goals of this Paper.} We provide a few definitions and
an example to indicate how the strand space framework can relate
choreographies to the cryptographic protocols that implement them.
In particular, we consider a very simple choreography language, and
provide a semantics for it as a set of ``abstract bundles.'' That is,
each session of the protocol executes according to one of the bundles
predicted by the semantics. Moreover, any collection of sessions that
may have occurred takes the following form: its events partition into
bundles that are obtained by instantiating the parameters in bundles
given in the semantics. Also, if two nodes belong to different
partition elements, there is no $\preceq$ ordering between them,
unless the executions are generated as parts of some higher-level
choreography that might determine a causal ordering.
We call this an {\em abstract bundle semantics} because it builds in
the assumptions of the choreography level: messages do not have
explicit cryptographic operations, and the choreography-level
communication assumptions are satisfied.
Messages are always delivered exactly once; sender and recipient are
never mismatched; no message is created by adversary operations. We
must connect this idealized semantics with a more realistic semantics
at the cryptographic level, in which the adversary may be active.
One peculiarity of our message datatype is that we allow ``boxes.'' A
box $\msgbox{\tilde{M}}{\rho\rho'}$ is a message prepared on role
$\rho$ that can be opened only by a principal playing role $\rho'$.
At the choreography level, this property is enforced by a type system.
We use these boxes to make explicit the confidentiality and
authentication requirements of a choreography in the case where some
roles are played by compromised participants. However, in this
article, we focus on the simplest case, in which no participants are
compromised. That is, we will assume here, that any participant who
is sent a box, will behave only as predicted by the choreography.
Our semantics at the \emph{cryptographic level} is a standard strand
space treatment, except for one ingredient. Namely, this semantics
assumes that some kinds of messages are delivered at most once. These
are session-initiating messages that contain a nonce, or in some
protocols a freshly generated session key. Implementations now use a
nonce-caching technique in which the nonces of previously executed
sessions are retained in a cache. A new incoming message contains a
nonce which is compared against the cache; if it is present, then with
overwhelming probability there has been a replay attempt, and the
message is discarded. Otherwise, the nonce is recorded and the
session proceeds. So as not to need to retain nonces forever,
implementations typically combine this with a timestamp, and assume
that uncompromised principals are loosely synchronized. A message
with too old a timestamp is discarded. Nonces may be dropped from the
cache when their timestamps have expired. In this approach, the nonce
and the timestamp must appear digitally signed in the incoming message
to prevent manipulation by the adversary.
We define a cryptographic protocol to properly implement a
choreography if, when abstracting its possible executions in this
at-most-once semantics, we obtain exactly the possible executions of
the abstract bundle semantics for the choreography.
We explore here a simple example in which the participants are
well-known to each other from the start of the transaction. However,
the ideas also apply when additional participants may be chosen during
execution, and keys must be distributed as part of the message flow.
\section{Strand Spaces}
\label{sec:strandspaces}
Strand spaces \cite{strandspaces,GuttmanEtAl05} were developed as a
simplest possible model for cryptographic protocol analysis, but are
also applicable to other kinds of distributed systems. In strand
spaces, we consider {\em strands}, behavioural traces for roles
represented as finite linear sequences of transmission and reception
events. The model provides techniques for analysing how various
strands can be combined together in a run of a protocol including some
adversary behaviour.
Let $A$ be a set of messages.
\begin{definition}[Strand Space]
A {\em directed term} is a pair denoted by $\pm a$ (for $a$ a
message $\in A$) where $\pm\in\{-,+\}$ is a direction with $+$
representing transmission and $-$ reception. A trace is an element
of $(\pm A)^*$, the set of infinite sequences of directed terms.
\noindent A {\em strand space} is a set $S$ equipped with a trace mapping
$\mathsf{tr}:S\rightarrow(\pm A)^*$ and its elements are called {\em
strands}.
\end{definition}
\noindent If $s$ is a strand in some strand space $S$ then its
$i^{\mathrm{th}}$ member denotes the $i^{\mathrm{th}}$ transmission or
reception event in $s$. Formally, we interpret this as the pair
$s,i$, which we call a \emph{node} on the strand $s$.
We write $m\Rightarrow n$ when, for some $s$ and $i$, $m=s,i$ and
$n=s,i+1$, i.e.~$n$ is the node immediately following $m$ on the
strand $s$. We write $\msg n$ for the message sent or received in the
directed term of $n$. That is, if $n=s,i$, and $s(i)$ is a
transmission $+t$ or reception $-t$ of message $t$, then $\msg n=t$.
We occasionally write $\dmsg n=\pm t$ for the message together with
its direction. We write $m\rightarrow n$ when for some $t$, $\dmsg
m=+t$ and $\dmsg n=-t$. Thus, $n$ could receive its message directly
from $m$.
\smallskip
But how can strands be combined together in order to represent
executions of a protocol? This is precisely captured by the notion of
{\em bundle} for a strand space $S$:
\begin{definition}[Bundle]
A finite acyclic directed graph
$\mathcal{B}=(\mathcal{N},\mathcal{E},\preceq_{\mathcal{B}})$ is a \emph{bundle}
for $S$ if
\begin{enumerate}
\item $\mathcal{N}$ is a set of strand nodes in $S$ such that if
$n\in\mathcal{N}$ and $m\Rightarrow n$, then $m\in\mathcal{N}$;
\item $\mathcal{E}=\rightarrow_{\mathcal{B}}\cup\Rightarrow_{\mathcal{B}}$ where
\begin{enumerate}
\item $\Rightarrow_{\mathcal{B}}$ is the restriction of $\Rightarrow$ to
nodes in $\mathcal{N}$;
\item $\rightarrow_{\mathcal{B}}\subseteq(\rightarrow\cap\mathcal{N} \times
\mathcal{N})$; and
\item for any reception node $n\in\mathcal{N}$, there is exactly
one transmission node $m\in\mathcal{N}$ such that
$m\rightarrow_{\mathcal{B}}n$.
\end{enumerate}
\end{enumerate}
$n\preceq_{\mathcal{B}} m$ iff there is a path using arrows
$\rightarrow_{\mathcal{B}}\cup\Rightarrow_{\mathcal{B}}$ from $n$ to $m$ in
$\mathcal{B}$.
\end{definition}
A \emph{bundle} is a causally well-founded graph -- essentially, a
Lamport diagram -- built from strands and transmission edges. The
relation $\preceq_{\mathcal{B}}$ is a well-founded partial order, meaning
that the \emph{bundle induction} principle holds, that every non-empty
set of nodes of $\mathcal{B}$ contains $\preceq_{\mathcal{B}}$-minimal members.
The notions of strand and bundle, and the principle of bundle
induction, are the essential ingredients in the strand space model.
Choices -- such as
what operations the adversary strands offer, or what additional
closure properties bundles may satisfy -- can vary to model different
problems concerning cryptographic protocols or distributed
communication more generally.
\smallskip
\noindent {\bf Example.} We briefly introduce an example in order to
clarify the concepts introduced above. Let $S$ be composed by the
following strands:
\begin{center}
(1) $n_1\Rightarrow n_2$\quad\qquad (2) $n_3\Rightarrow
n_4$\quad\qquad (3) $n_5\Rightarrow n_6$\quad\qquad (4)
$n_7\Rightarrow n_8\Rightarrow n_9\Rightarrow n_{10}$\quad\qquad
(5) $n_{11}\Rightarrow n_{12}$
\end{center}
where
{\small
\begin{center}
\begin{tabular}{llll}
$\mathsf{dmsg}(n_1)=+\text{``Hello''}$ &
$\mathsf{dmsg}(n_2)=-\text{``Bye''}$ \\
$\mathsf{dmsg}(n_3)=+\text{``Good\ luck''}$ &
$\mathsf{dmsg}(n_4)=-\text{``Thanks''}$ \\
$\mathsf{dmsg}(n_5)=-\text{``Good\ luck''}$ &
$\mathsf{dmsg}(n_6)=+\text{``Thanks''}$ \\
$\mathsf{dmsg}(n_7)=-\text{``Hello''}$ &
$\mathsf{dmsg}(n_8)=-\text{``Good\ luck''}$ &
$\mathsf{dmsg}(n_9)=+\text{``Thanks''}$ &
$\mathsf{dmsg}(n_{10})=+\text{``Bye''}$ \\
$\mathsf{dmsg}(n_{11})=-\text{``Thanks''}$ &
$\mathsf{dmsg}(n_{11})=+\text{``Bye''}$
\end{tabular}
\end{center}
}
Below, we report two possible executions in the strand space $S$
(for clarity, we label $\rightarrow$ with the corresponding message):\\[1mm]
\begin{center}
\begin{tabular}{l}
\begin{diagram}
n_3 & \rTo^{\quad \text{Good luck}\quad} & n_5\\
\dImplies & & \dImplies\\\\
n_4 & \lTo^{\quad\text{Thanks}\quad} & n_6
\end{diagram}
\end{tabular}
\qquad
\begin{tabular}{l}
\begin{diagram}
n_1 & \rTo^{\quad\text{``Hello''}\quad} & n_7\\
\dImplies & & \dImplies\\\\
& & n_8 & \lTo^{\quad\text{``Good Luck''}\quad} & n_3\\
& & \dImplies & & \dImplies\\\\
& & n_9 & \rTo^{\quad\text{``Thanks''}\quad} & n_4\\
& & \dImplies \\\\\\
n_2 & \lTo^{\quad\text{``Bye''}\quad} & n_{10}
\end{diagram}
\end{tabular}
\end{center}
Note that strand (5) could interfere allowing for the following
bundle:
\begin{center}
\begin{tabular}{l}
\begin{diagram}
n_1 & \rTo^{\quad\text{``Hello''}\quad} & n_7\\
\dImplies & & \dImplies\\\\
& & n_8 & \lTo^{\quad\text{``Good Luck''}\quad} & n_3\qquad\\
& & \dImplies\\\\
& & n_9 & \rTo^{\quad\text{``Thanks''}\quad} & & &n_{11}\\
& & & & & &\dImplies\\\\\\
n_2 & \lTo^{\quad\text{``Bye''}\quad} & & & & &n_{12}
\end{diagram}
\end{tabular}
\end{center}
\section{An Execution Model for Choreography}
\label{sec:choreography}
\subsection{The Calculus}
\paragraph{Syntax.} Let $\rho$ range over the set of roles $\mathcal
R$. The syntax of our choreography mini-language (based on the Global
Calculus \cite{carbone.honda.yoshida:esop07}) is given by the
following grammar:
\begin{align*}
C::=&\phantom{{}\mid\quad{}}\Sigma_i\,\interact{\rho_1}{\rho_2}{op_i}{\tilde
M_i}\mathbf.\ C_i\mid\quad \mathbf0 \qquad&\qquad M::= &
\phantom{{}\mid\quad{}}v\mid \msgbox{\tilde
M}{\rho_1\rho_2}
\end{align*}
\noindent Above, the term $\Sigma_i\,\interact{\rho_1}{\rho_2}{op_i}{\tilde
M_i}\mathbf.\ C_i$ describes an interaction where a branch with label
${\sf op}_i$ is non-deterministically selected and a message $\tilde
M_i$ is sent from role $\rho_1$ to role $\rho_2$. Each two roles in a
choreography share a private channel hence it would be redundant to
have them explicit in the syntax \cite{BCDDDY:concur2008}.
Term $\mathbf0 $ denotes the inactive system. A message $M$ can either be
a value $v$ or a box $\msgbox{\tilde M}{\rho_1\rho_2}$. The latter
denotes a tuple of messages $M_i$ from $\rho_1$ that can only be
opened by $\rho_2$.
\paragraph{Syntactic Assumption.} The \emph{sender} of a choreography
of the form $\Sigma_i\,\interact{\rho_1}{\rho_2}{op_i}{\tilde M_i}\mathbf.\
C_i$ is $\rho_1$. We assume, for every choreography $C$:
\begin{itemize}
\item all $\op{op}$'s are distinct.
\item in any path in a choreography syntax tree, a box
$\msgbox{\tilde M}{\rho_1\rho_2}$ has to occur first in an
interaction whose sender is $\rho_1$ and can only be opened by
$\rho_2$ in later interaction;
\item if $C=\Sigma_i\,\interact{\rho_1}{\rho_2}{op_i}{\tilde
M_i}\mathbf.\ C_i$ then either $C_i=\mathbf0 $ or the sender of $C_i$ is
$\rho_2$ for all $C_i$;
\end{itemize}
The last assumption above requires that the receiving role in an
interaction is always the transmitting role in the subsequent
interaction. All the assumptions above can be statically
checked~\cite{CG09b}.
\paragraph{LTS Semantics.} Our mini-language can be equipped with a
standard trace semantics with configurations $C\lts{\mu} C'$ where
$\mu$ contains the parameters of the interaction performed i.e. it
ranges over the set $\mathcal R\times\mathcal R\times\mathcal
{O}\times \mathcal M$ where $\mathcal O$ is the set of operators
$\op{op}$ and $\mathcal M$ the set of messages. The following rule
formally defines the relation $\lts{\mu}$ which is taken up to
commutativity and associativity of $+$:
\begin{align*}
\Did{C-Com}\
&\
\Rule
{}
{
\Sigma_i\,\interact{\rho_1}{\rho_2}{op_i}{\tilde M_i}\mathbf.\ C_i
\quad\lts{(\rho_1,\rho_2,\op{op_i},\tilde M)}\quad
C_i
}
\end{align*}
\paragraph{Buyer-Seller Example.}
We report a variant of the Buyer-Seller financial protocol
\cite{carbone.honda.yoshida:esop07}. A buyer (or client) $\mathsf{C}$
asks a seller $\mathsf{S}$ for a quote about a product
$\mathsf{prod}$. If the quote is accepted, $\mathsf{C}$ will send its
credit card $\mathsf{card}$ to $\mathsf{S}$ who will forward it to a
bank $\mathsf{B}$. The
bank
will check if the payment can be done and, if so, reply with a receipt
$\mathsf{receipt}$ which $\mathsf S$ will forward to $\mathsf{C}$. In
our syntax:
\begin{align*}
1.\quad &
\interact{\mathsf{C}}{\mathsf{S}}{\texttt{req}}{\mathsf{prod}}\mathbf.\ \
\interact{\mathsf{S}}{\mathsf{C}}{\texttt{reply}}{\mathsf{quote}}\mathbf.\ \\
2.\quad & (\quad\interact{\mathsf{C}}{\mathsf{S}}{\texttt{ok}}
{\msgbox{\mathsf{card}}{\mathsf{C}\mathsf{B}}}\mathbf.\ \
\interact{\mathsf{S}}{\mathsf{B}}{\texttt{pay}}
{\msgbox{\mathsf{card}}{\mathsf{C}\mathsf{B}}}\mathbf.\
(\quad\interact{\mathsf{B}}{\mathsf{S}}{\texttt{okcf}}
{\msgbox{\mathsf{receipt}}{\mathsf{B}\mathsf{C}}}\mathbf.\ \\
3.\quad &
\phantom{
(\quad\interact{\mathsf{C}}{\mathsf{S}}{\texttt{ok}}
{\msgbox{\mathsf{card}}{\mathsf{C}\mathsf{B}}}\mathbf.\ \
\interact{\mathsf{S}}{\mathsf{B}}{\texttt{pay}}
{\msgbox{\mathsf{card}}{\mathsf{C}\mathsf{B}}}\mathbf.\
(\quad
}
\interact{\mathsf{S}}{\mathsf{C}}{\texttt{rcpt}}
{\msgbox{\mathsf{receipt}}{\mathsf{B}\mathsf{C}}}\\
4.\quad &
\phantom{
(\quad\interact{\mathsf{C}}{\mathsf{S}}{\texttt{ok}}
{\msgbox{\mathsf{card}}{\mathsf{C}\mathsf{B}}}\mathbf.\ \
\interact{\mathsf{S}}{\mathsf{B}}{\texttt{pay}}
{\msgbox{\mathsf{card}}{\mathsf{C}\mathsf{B}}}\mathbf.\
(\quad
}
\quad\qquad\qquad+\\
5.\quad &
\phantom{
(\quad\interact{\mathsf{C}}{\mathsf{S}}{\texttt{ok}}
{\msgbox{\mathsf{card}}{\mathsf{C}\mathsf{B}}}\mathbf.\ \
\interact{\mathsf{S}}{\mathsf{B}}{\texttt{pay}}
{\msgbox{\mathsf{card}}{\mathsf{C}\mathsf{B}}}\mathbf.\
(\quad
}
\interact{\mathsf{B}}{\mathsf{S}}{\texttt{nopaycf}}
}\mathbf.\ \\
6.\quad &
\phantom{
(\quad\interact{\mathsf{C}}{\mathsf{S}}{\texttt{ok}}
{\msgbox{\mathsf{card}}{\mathsf{C}\mathsf{B}}}\mathbf.\ \
\interact{\mathsf{S}}{\mathsf{B}}{\texttt{pay}}
{\msgbox{\mathsf{card}}{\mathsf{C}\mathsf{B}}}\mathbf.\
(\quad
}
\interact{\mathsf{S}}{\mathsf{C}}{\texttt{nopay}}
}\quad)\\
7.\quad & \qquad\qquad\qquad\qquad+\\
8.\quad & \phantom{(\quad}
\interact{\mathsf{C}}{\mathsf{S}}{\texttt{refuse}}{\mathsf{reason}})
\end{align*}
\noindent Line 1. denotes the quote request and reply. Lines 2. and 8. are
computational branches corresponding to acceptance and rejection of
the quote respectively. If the quote is accepted, $\mathsf{C}$ will
send its credit card in the box
$\msgbox{\mathsf{card}}{\mathsf{C}\mathsf{B}}$ meaning that
$\mathsf{S}$ cannot see it. The card number is then forwarded to
$\mathsf{B}$ who can open the box (line 2.). If the transaction can be
finalised a receipt is forwarded to $\mathsf{C}$. Otherwise, a
\texttt{nopay} notification will be sent. $\mathsf{B}$ boxes the
receipt so that it cannot be seen or changed by $\mathsf{S}$.
\subsection{Abstract Bundle Semantics (ABS).}
We introduce an alternative semantics for choreography based on
bundles defined as judgements of the form:
\[\semproves{} C{ \{(\mathcal B_1,\mathsf{who}_1),\ldots,(\mathcal
B_i,\mathsf{who}_i)\}}\] where $(\mathcal B,\mathsf{who}{})$ is a {\em bundle
environment}. Given a role $\rho$, $\mathsf{who}(\rho)$ denotes the strand
in the bundle $\mathcal B$ associated to the behaviour of $\rho$. The
abstract bundle semantics $\sembrack{C} = { \{(\mathcal
B_1,\mathsf{who}_1),\ldots,(\mathcal B_i,\mathsf{who}_i)\}}$ if and only if
$\semproves{} C{ \{(\mathcal B_1,\mathsf{who}_1),\ldots,(\mathcal
B_i,\mathsf{who}_i)\}}$. The relation $\models$ is the minimum relation
satisfying the following:
\begin{displaymath}
\begin{array}{rl}
\Did{ABS-Com}\
&
\Rule
{
\begin{array}{l}
\forall i\mathbf.\ \semproves{}{C_i}{\{(\mathcal B_{i1},\mathsf{who}_{i1}),\ldots,(\mathcal B_{ij_i},\mathsf{who}_{ij_i})\}}
\end{array}
}
{
\semproves{}
{\Sigma_i\,\interact{\rho_1}{\rho_2}{op_i}{\tilde M_i}\mathbf.\ C_i}
{
\left(
\begin{array}{c}
\bigcup_i \{(\mathcal B_{ij_i},\mathsf{who}_{ij_i})\}_{j_i}[\rho_1,\rho_2,\op{op}_i(\tilde M_i)]\\
\end{array}
\right)
}
}
\end{array}
\end{displaymath}
\begin{displaymath}
\begin{array}{rl}
\Did{ABS-Zero}\
&
\Rule
{
\op e\text{ fresh}
}
{
\semproves{\emptyset}{\mathbf0 }{(\{\op e^\rho\}_\rho,\lambda \rho\mathbf.\ \op e^\rho)}
}
\end{array}
\end{displaymath}
\noindent The abstract bundle semantics provides a set of bundles which
represents all executions of the protocol described by the
choreography. In \Did{ABS-Com}, $(\mathcal
B_{ij_i},\mathsf{who}_{ij_i})[\rho_1,\rho_2,\op{op}_i(\tilde M_i)]$ denotes a
new bundle obtained from $\mathcal B_{ij_i}$ where the two strands
$\mathsf{who}_{ij_i}(\rho_1)$ and $\mathsf{who}_{ij_i}(\rho_2)$ are prefixed with the
events $+\op{op}_i(\tilde M_i)$ and $-\op{op}_i(\tilde M_i)$
respectively. The function $\mathsf{who}_{ij_i}$ is updated
accordingly. Formally,
\begin{equation*}
(\mathcal B,\mathsf{who})[\mu]\ = \
(\quad
(
\mathcal N\cup\{n_i\}_i,
\mathcal E\cup\{n_i\Rightarrow \mathsf{who}(\rho_i)\}_{i}\cup\{n_1\rightarrow n_2\},
\preceq'
),\quad
\mathsf{who}[\rho_i\mapsto n_i\Rightarrow \mathsf{who}(\rho_i)]_i
\quad)
\end{equation*}
where $\prec'$ is the update of $\prec_{\mathcal B}$ according to the
new elements added to the bundle and $\mathcal B=(\mathcal N,\mathcal
E,\preceq_{\mathcal B})$. The operation above is applied to all those
bundles obtained from the semantics of each branch and the result will
be their union. In \Did{ABS-Zero}, we augment the set $A$ with fresh
events $\{\op e^\rho\}\in E$ in order to distinguish each strand.
\paragraph{ABS Example.} The ABS for the Buyer-Seller protocol has
three bundles corresponding to its possible executions, namely: (i)
$\mathsf{C}$ accepts the quote and $\mathsf{B}$ successfully finalises
the transaction sending back a receipt; (ii) $\mathsf{C}$ accepts the
quote but $\mathsf{B}$ does not accept the payment; and (iii)
$\mathsf{Buyer}$ does not accept the quote with reason
$\mathsf{reason}$ and the protocol terminates. The three corresponding
bundles are reported in Fig.~\ref{fig:exabs}.
\begin{figure}
\centering
\begin{tabular}{lll}
(i)\quad
\begin{diagram}
\mathsf{C} & \rTo^{\texttt{req}\langle\mathsf{prod}\rangle} & \mathsf{S} \\\\\\
\dImplies & & \dImplies & & \\
\bullet & \lTo^{\texttt{reply}\langle\mathsf{quote}\rangle} & \bullet \\\\\\
\dImplies & & \dImplies & & \\
\ast &
\rTo^{\texttt{Ok}\langle\msgbox{\mathsf{card}}{\mathsf{C}\mathsf{B}}\rangle}&
\bullet &
\rTo^{\texttt{pay}\langle\msgbox{\mathsf{card}}{\mathsf{C}\mathsf{B}}\rangle}
& \mathsf{B} \\\\\\
\dImplies&&\dImplies&&\dImplies\\
\bullet &
\lTo^{\texttt{rcpt}\langle\msgbox{\mathsf{receipt}}{\mathsf{B}\mathsf{C}}\rangle}
&\bullet&
\lTo^{\texttt{okcf}\langle\msgbox{\mathsf{receipt}}{\mathsf{B}\mathsf{C}}\rangle}
&\ast
\end{diagram}
\end{tabular}
\\[5mm]
\begin{tabular}{lll}
(ii)\quad
\begin{diagram}
\mathsf{C} &
\ \ldots\text{as in (i)}\ldots\ &
\mathsf{S} &
\ \ldots\text{as in (i)}\ldots\ &
\mathsf{B}
\\\\\\
\dImplies&&\dImplies&&\dImplies\\\\\\
\bullet&\lTo^{\texttt{nopay}\langle\rangle}
&\bullet&\lTo^{\texttt{nopaycf}\langle\rangle}
&\ast
\end{diagram}
\quad\qquad&\quad\qquad
(iii)\quad
\begin{diagram}
\mathsf{C} & \rTo^{\texttt{req}\langle\texttt{prod}\rangle} & \mathsf{S} \\\\\\
\dImplies & & \dImplies & & \\
\bullet & \lTo^{\texttt{reply}\langle\mathsf{quote}\rangle} & \bullet \\\\\\
\dImplies & & \dImplies & & \\
\ast & \rTo^{\texttt{refuse}\langle\mathsf{reason}\rangle} & \bullet
\end{diagram}
\end{tabular}
\caption{Bundles for the Buyer-Seller protocol}
\label{fig:exabs}
\end{figure}
The nodes marked with $\ast$ are those points where there is a
possibility of branching i.e. bundle (ii) is identical to (i) up to
its $\ast$ while (iii) is identical to (i) and (ii) up to its
$\ast$. Note that (iii) only involves roles $\mathsf C$ and $\mathsf
S$.
\smallskip
\smallskip
\noindent In the sequel, let $(\mathcal B,\mathsf{who})\backslash[\mu]$ be defined
as follows:
\begin{equation*}
(\mathcal B,\mathsf{who})\backslash[\mu]\ = \
\left\{
\begin{array}{ll}
\mathcal B' & \text{ if}\quad\mathcal B = (\mathcal B',\mathsf{who})[\mu]\\
\text{undefined} & \text{ otherwise}
\end{array}
\right.
\end{equation*}
Intuitively, the operation above is inverse to $(\mathcal
B,\mathsf{who})[\mu]$ i.e. removes the first communication from a bundle (if
equal to $\mu$, undefined otherwise). We can then conclude this
section with a result that relates the LTS semantics to the bundle
semantics.
\begin{theorem}\label{theorem}
Let $C$ be a choreography. Then,
\begin{enumerate}
\item if $C\lts{\mu}C'$ then there exists a bundle $\mathcal B$ in
$\sembrack{C}$ such that
$\sembrack{C'}=\sembrack{C}\backslash(\{\mathcal B\}\cup
L)\cup\{\mathcal B\backslash[\mu]\}$ for $L = \{\mathcal B'\mathrel{\boldsymbol{\mathord{\mid}}}
\mathcal B\in\sembrack{C}\land\mathcal B\backslash[\mu]\text{ is
undefined}\}$;
\item if $\mathcal B\backslash[\mu]$ is defined and $\mathcal
B\in\sembrack{C}$ then there exists $C'$ such that $C\lts\mu C'$.
\end{enumerate}
\end{theorem}
\section{An execution model for Cryptoprotocols}
\label{sec:cryptoprotocols}
Cryptographic protocols are modelled by strand spaces where the set of
messages $a$ is more general. Formally, crypto-level messages, denoted
by the syntactic category $t$ have the following syntax:
\begin{align*}
t::= & \phantom{{}\mid\quad{}}\tilde v\quad\mid\quad \enc{\tilde
t}{K}
\end{align*}
Above, the value $v$ ranges over the disjoint union of infinite sets
of nonces (denoted by $N$), atomic keys (denoted by $K$) and other
basic values. We will write a sequence of messages in the form
$v_1,\;\;\ldots,\;\; v_k$. A node of a protocol $\Pi$ is
\emph{regular} if it lies on a strand of $\Pi$, not on an adversary
strand.
\begin{definition}[Deliver-once] Suppose that $S$ is a set of
messages, and $\mathcal{B}$ is a bundle. $\mathcal{B}$ \emph{delivers messages
in} $S$ \emph{only once} if there exists an injective
function $f\colon R\rightarrow T$, where
\begin{itemize}
\item $R$ is the set of regular nodes $n$ in $\mathcal{B}$ such that a
member of $S$ is received on $n$, and
\item $T$ is the set of regular nodes $n$ in $\mathcal{B}$ such that a
member of $S$ is transmitted on $n$.
\end{itemize}
When $\{S_i\}_{i\in I}$ is a family of sets indexed by $i\in I$, we
say that $\mathcal{B}$ is \emph{deliver-once} for $\{S_i\}_{i\in I}$ when
$\mathcal{B}$ delivers messages in each $S_i$ only once.
\end{definition}
We typically apply this definition when $I$ is a set of values that
will be generated freshly, and $S_i$ is a set of messages of
particular forms containing one such value $i$ ($K_{j,k}$ in the
example below).
\paragraph{Cryptoprotocol Example.} The Buyer-Seller cryptoprotocol
implements the choreography example of Section~\ref{sec:choreography}.
It provides parametric strands that define the behaviors of the
principals as they send and receive encrypted messages to provide
security services for the behaviors in the choreography. The central
idea is that the first few messages use public encryption keys and
nonces to establish symmetric keys. The remaining messages then use
the keys in a straightforward way. To establish a key between $A$ and
$B$, $A$ sends a message containing a nonce, encrypted with $B$'s
public key. $B$ returns a message encrypted with $A$'s public key.
It contains $A$'s nonce as well as a fresh symmetric key to be used
for this session. We use different syntactic tags in each encrypted
unit which correspond to the $\op{op}$'s in the choreography (denoted
by the typewriter font \texttt{op}). At this level, the tags ensure
that no unit can be confused with any other (this is the reason why
the $\op{op}$'s are all distinct at choreography level).
The key exchange phase takes the form shown in
Fig.~\ref{fig:key:exchange}.
\begin{figure}
\centering
\begin{diagram}[h=3mm,w=8mm]
C & \rTo{m_1,\;\; m_2} & \qquad & \rTo{m_1,\;\; x} & S &&&& \\
\dStrNext & & & & \dStrNext & & & & \\
&&&& \bullet & \rTo{m_3,\;\; x} & \qquad &\rTo{m_3,\;\; m_2} & B\\
&&&& \dStrNext & & & & \dStrNext\\
&&&& \bullet & \lTo{m_4,\;\; y}&\qquad&\lTo{m_4,\;\;
m_5}&\bullet \\
&&&& \dStrNext &&&& \\
\bullet& \lTo{m_6,\;\; m_5} &\qquad&\lTo{m_6,\;\; y} & \bullet
&&&& \\
\dStrNext &&&& \dStrNext &&&& \dStrNext \\
\null &&&& \null &&&& \null
\end{diagram}
\begin{tabular}[c]{l@{\qquad\qquad}r}
$ m_1 = \enc{\texttt{cs}\; C,\;\; B,\;\; N_1}{\pk{S}}$ &
$ m_2 = \enc{\texttt{cb}\; C,\;\; S,\;\; N_1}{\pk{B}}$ \\
$ m_3 = \enc{\texttt{sb}\; C,\;\; S,\;\; N_1}{\pk{B}}$ &
$ m_4 = \enc{\texttt{bsk}\; N_2,\;\; N_1,\;\; K_{bs}}{\pk{B}}$ \\
$ m_5 = \enc{\texttt{bck}\; N_1,\;\; K_{bc}}{\pk{C}}$ &
$ m_6 = \enc{\texttt{sck}\; N_1,\;\; N_2,\;\; K_{sc}}{\pk{C}}$
\end{tabular}
\caption{Key exchange phase}
\label{fig:key:exchange}
\end{figure}
Each participant leaves the key exchange phase knowing that $N_1,N_2$
are shared among $C,S,B$, and that two symmetric keys are to be used
for encryption in the next phase. For instance, $C$ knows to use
$K_{sc}$ to communicate with the seller in the ensuing exchange, and
to use $K_{bc}$ to communicate with the bank.
In the ensuing stage, the participants use these keys to transfer the
payloads amongst themselves. Their exchange---in the successful case,
in which the transaction completes---takes the form shown in
Fig.~\ref{fig:payloads}.
\begin{figure}[th]
\centering
\begin{diagram}[h=3mm,w=8mm]
C & \rTo{p_1} & \qquad & \rTo{p_1} & S &&&& \\
\dStrNext & & & & \dStrNext & & & & \\
\bullet & \lTo{p_2} & \qquad & \lTo{p_2} & \bullet &&&& \\
\dStrNext & & & & \dStrNext & & & & \\
\ast & \rTo{p_3} & \qquad & \rTo{p_3} & \bullet &&&& \\
\dStrNext & & & & \dStrNext & & & & \\
&&&& \bullet & \rTo{p_4} & \qquad &\rTo{p_4} & B\\
&&&& \dStrNext & & & & \dStrNext\\
&&&& \bullet & \lTo{p_5}&\qquad&\lTo{p_5[p_6/y]} & \ast \\
&&&& \dStrNext &&&& \\
\bullet& \lTo{p_6} &\qquad&\lTo{y} & \bullet
&&&&
\end{diagram}
\ \\
\ \\
\ \\
\begin{tabular}[c]{l@{\qquad\qquad}r}
$ p_1 = \enc{\texttt{req}\; N_2 ,\;\; C,\;\; S,\;\; B,\;\; \mathsf{prod}}{K_{sc}}$ &
$ p_2 = \enc{\texttt{reply}\; \mathsf{quote}}{K_{sc}}$ \\
$ p_3 = \enc{\texttt{ok}\; x}{K_{sc}}$ &
$ p_4 = \enc{\texttt{pay}\; x ,\;\; C,\;\; S}{K_{bs}}$ \\
$ p_5 = \enc{\texttt{okcf}\; y}{K_{bs}}$ &
$ p_6 = \enc{\texttt{rcpt}\; \mathsf{receipt}}{K_{bc}}$
\end{tabular}
\caption{Payload exchange phase}
\label{fig:payloads}
\end{figure}
However, $\mathsf{C}$ and $B$ each have an opportunity to prevent the
exchange from completing, at the nodes marked $\ast$. If $C$
transmits $\enc{\texttt{refuse}}{K_{sc}}$ instead of $p_3$, then $S$
must terminate the exchange before contacting $B$. If $B$ transmits
$\enc{\texttt{nopaycf}\; \enc{\texttt{nopay}}{K_{bc}}}{K_{bs}}$ instead
of $p_5[p_6/y]$, then $S$ and $C$ must terminate the transaction.
Let us assume that the participants of a run use their private
decryption keys only in accordance with this protocol, and that the
nonces $N_1,N_2$ and keys $K_{bc},K_{bs},K_{sc}$ are in fact freshly
chosen and unguessable. On this assumption, there are essentially
only three possible executions, if we consider only those of minimal
size, given that a role completed. When $C$ completes normally, then
the other participants have completed normally with matching
parameters. When $S$ completes with a client refusal, then $C$ has
refused and $B$ has had a matching key exchange phase but no more.
When $C$ completes with a $\texttt{nopay}$ message, then $B$ has
refused to pay, and $S$ has been informed of this. This analysis
indicates that the protocol appears to achieve its goals. Indeed, we
have confirmed this with the tool \textsc{cpsa}, a Cryptographic
Protocol Shapes Analyzer~\cite{DoghmiGuttmanThayer07}, which
enumerates the minimal, essentially different executions of the
protocol. We can then check the assertions we have just made by
inspecting those executions.
\section{Abstraction and Correctness} A partial function $\alpha$ over
messages is an \emph{abstraction map} if (1) $\alpha(t)$ (if defined)
contains no cryptographic operators, nonces nor keys, and (2) the
parameters in $\alpha(t)$ (if defined) always appear in $t$.
For instance, $\alpha$ could map $\enc{\texttt{req}\; N_2 ,\;\; C,\;\;
S,\;\; B,\;\; \mathsf{prod}}{K_{sc}}$ to
${\texttt{req}\langle\mathsf{prod}\rangle}$ in our Buyer-Seller
example. The result has no cryptography and no nonces, and the tags
$\texttt{req}$ and $\mathsf{prod}$ appear in the argument.
We say that an abstract strand $s$ is an \emph{image} of a
cryptographic strand $s_c$ if, ignoring transmissions or receptions on
$s_c$, for which $\alpha$ is undefined, for each transmission or
reception node $n$ on $s$, its message $\kind{msg\/}(n)$ is
$\alpha(\kind{msg\/}(n_c))$, where $n_c$ is the corresponding transmission or
reception node (resp) on $s_c$. That is, $\alpha$ yielding the trace
of $s$, when mapped through the trace of $s_c$ restricted to the
domain of $\alpha$.
Suppose that a concrete strand $s_c$ has its first $i$ nodes in a
concrete bundle $\mathcal{C}$, but $\alpha$ is undefined for the messages on
these nodes. We then say that $s_c$ is \emph{abstractly vacuous in}
$\mathcal{C}$. In the opposite case, when some node $n$ of $s_c$ is in
$\mathcal{C}$ and $\alpha(\kind{msg\/}(n))$ is well-defined, we say that $s_c$ is
\emph{abstractly non-vacuous in} $\mathcal{C}$.
An abstract bundle $\mathcal{B}$ is an \emph{image} of a cryptographic bundle
$\mathcal{C}$ if (1) there is a bijection $\phi$ between the abstractly
non-vacuous regular strands $s_c$ of $\mathcal{C}$ and the regular strands
$s$ of $\mathcal{B}$; (2) $\phi(s_c)$ is always an image of $s_c$; and (3)
the transmission relation $\rightarrow_{\mathcal{B}}$ is formed by connecting
nodes of $\mathcal{B}$ such that $m\rightarrow_{\mathcal{B}}n$ implies
$m_c\preceq_{\mathcal{C}}n_c$, for some concrete nodes of which $m,n$ are
images. See~\cite{Guttman09a} for a related notion of protocol
transformation, and~\cite{MaffeiEtAl07} for an approach to protocol
verification via abstraction functions.
Suppose that $\mathcal{C}$ is a concrete bundle and $\{\mathcal{C}_i\}_i$ is a
family of sub-graphs of $\mathcal{C}$ that partitions the regular nodes of
$\mathcal{C}$. We say that $\{\mathcal{C}_i\}_i$ \emph{separates} $\mathcal{C}$
\emph{into components} when each $\mathcal{C}_i$ is a bundle on its own.
\begin{definition}[Faithfulness] \label{def:faithful}
Cryptoprotocol $\Pi$ is \emph{faithful to} choreography $C$ if there
is an abstraction function $\alpha$ such that:
\begin{enumerate}
\item Every $\mathcal{B}\in\sembrack{C}$ is an image of some bundle $\mathcal{C}$
of $\Pi$;\label{clause:fth:upward:cover}
\item If $\mathcal{C}$ is a bundle of $\Pi$, then some family
$\{\mathcal{C}_i\}_i$ separates $\mathcal{C}$ into components. Moreover, each
image $\mathcal{B}_i$ of any $\mathcal{C}_i$ is an initial sub-bundle of
$\sigma(\mathcal{B})$, for some $\mathcal{B}\in\sembrack{C}$ and some substitution
$\sigma$.\label{clause:fth:downward:separate}
\end{enumerate}
If $\{S_i\}_{i\in I}$ is a family of sets of messages, then $\Pi$ is
\emph{faithful to} $C$ \emph{assuming the deliver-once property for}
$\{S_i\}_{i\in I}$ if the above holds for bundles of $\Pi$ that are
deliver-once for $\{S_i\}_{i\in I}$.
\end{definition}
\smallskip
\noindent {\bf Faithfulness in the Buyer-Seller protocol.} We use the
protocol analysis tool \textsc{cpsa}~\cite{DoghmiGuttmanThayer07} as
part of a proof that the protocol of Fig.~\ref{fig:key:exchange} and
Fig.~\ref{fig:payloads} is faithful to the choreography in
Fig.~\ref{fig:exabs}. There are three stages:
\begin{enumerate}
\item \textsc{cpsa} determines the minimal, essentially different
executions that are possible, given that any one party has had a
complete run.
These are the expected success execution $_s$ and failure
execution $_f, _{f'}$, modulo the fact that a party never
knows whether its last message was successfully delivered, if its
last action is a transmission. In particular, the active parties
agree on all parameters to the session.
\item Based on this \textsc{cpsa} output, inspection shows that
Def.~\ref{def:faithful}, Clause~\ref{clause:fth:upward:cover} is
satisfied: Any run $\mathcal{B}\in\sembrack{C}$ is the abstraction of some
concrete bundle $\mathcal{C}$.
\item Because $_s,_f,_{f'}$ are the only minimal
forms of execution, every larger execution $\mathcal{B}_c$ is a (possibly
non-disjoint) union of executions of these forms. That is, there is
a family of maps $\{H_i\}_i$, where each $H_i$ maps either $_s$
or $_f$ to some subset of the regular nodes of $\mathcal{B}_c$.
Moreover, each regular node $n\in\mathcal{B}_c$ is the image of some node
in $_s,_f$, or $_{f'}$ under at least one of the
$H_i$.
However, each pair of strands agrees on a pair of freshly chosen
values, where each of them has chosen one of the values. This
forces the range of $H_i$ and $H_j$ either to coincide or be
disjoint. Hence Clause~\ref{clause:fth:downward:separate} is
satisfied when we define the family $\{\mathcal{C}_i\}_i$ by saying that
two nodes belong to the same $\mathcal{C}_i$ if they are both in the range
of any one $H_i$.
\end{enumerate}
\section{Concluding Remarks}
We have introduced two execution models, one for choreography
(assuming no compromised participants) and one for cryptoprotocols
with deliver-once assumptions. The abstract bundle semantics gives a
set of bundles representing all the possible runs of the protocol
described by a choreography. We have sketched a form of argument for
proving that a cryptoprotocol is faithful to the ABS of a
choreography.
In \cite{CG09b}, we studied an abstract semantics for the choreography
language presented here where roles can belong to compromised
principals. The ideas of abstraction have yet to be extended to the
compromised case and to a choreography language with infinite
states.
The work by Bhargavan et al. in \cite{BhargavanEtAl09, CDFBL08} is
closely related to ours: they provide a compiler for generating ML
code that can then be type-checked for verifying its security
property. Their notion of faithfulness is guaranteed for the
well-typed code generated from the source choreography.
In future work, we aim at developing systematic techniques for proving
that certain transformations preserve all of the goals of a protocol,
while achieving additional goals~\cite{Guttman09a}.
\label{sect:bib}
\bibliographystyle{plain}
|
\section{Experiments and Analysis}
Using high resolution detector telescopes with excellent isotope
identification capabilities, we recently studied a number of heavy
ion reactions to determine relative yields for production of a wide
range of isotopes~\cite{chen,huang10}.
The experiment was performed at the K-500 superconducting cyclotron
facility at Texas A$\&$M University. 40 A MeV $^{64}$Zn,$^{70}$Zn and
$^{64}$Ni beams irradiated $^{58}$Ni,$^{64}$Ni, $^{112}$Sn,$^{124}$Sn,
$^{197}$Au, and $^{232}$Th targets.
Intermediate mass fragments (IMFs) were detected by a detector telescope
placed at 20 degrees relative to the beam direction. The telescope
consisted of four Si detectors. Each Si detector had an area of 5cm
$\times$ 5cm.
The thicknesses are 129, 300, 1000 and 1000 $\mu$m.
Using the
$\Delta E-E$ technique we were typically able to identify 6-8 isotopes
for a given Z up to Z=18 with energy thresholds of 4-10 A MeV.
More details of the analysis are contained in ref.~\cite{chen,huang10}.
Isoscaling analyses were carried out for all possible combinations of
these reactions. Eighteen different reactions are considered here and therefore
more than 150 combinations are studied.
Data for each atomic number were independently fit to extract the
isoscaling parameter $\alpha(Z)$. $\beta(N)$ values were also extracted
for each neutron number. For some systems the extracted $\alpha(Z)$ parameter shows a steady decrease as Z increases. The $\beta(N)$ parameter generally showsa much smaller variation with increasing N, and has the opposite sign. A clear correlation between them, i.e. $\alpha(Z) \sim -\beta(N)$ for the equivalent number of nucleons, $N = Z$, has also been observed as suggested in the introduction(see Eq.(\ref{eq:invarian})). In Fig.1, the extracted isoscaling parameters for the case of Z = N = 7 are shown as a typical example.
Similar correlations are also observed for other selections of Z and N values if Z = N.
\begin{figure}[ht]
\includegraphics[width=3.6in,clip]
{Isoscaling_BetaVsAlpha_givenZN_groupZ7_fig1.eps}
\caption{\label{fig:fig_1}(a) $\beta(N)$ vs $\alpha(Z)$ is plotted for N=Z=7. Line indicates
the locus $\beta(N) =- \alpha(Z)$;
(b)ratio of $\beta(N)/\alpha(Z)$ data to theoretical Eq.(\ref{eq:kai}) (open squares);
(c)$\beta(N)/\alpha(Z)$ vs. $\alpha(Z)$ from data (open triangles);
(d)the analytical prediction Eq.(\ref{eq:kai}) (open circles) to compare with data. }
\end{figure}
In part a of Figure 1 we have plotted $\beta(N)$ vs $\alpha(Z)$.
As seen in the figure, the relation $\alpha(Z) \sim -\beta(N)$ is observed
for $\alpha(Z) \le 0.5$
and may deviate slightly at the larger $\alpha(Z)$ values.
Those larger values are associated with the largest N/A values
for the compound system. In the bottom part of the figure, these values on the left are
compared to predictions of Eq.(\ref{eq:kai}) on the right with the assumptions that Z/A
is that of the compound system.
We note that the experimental values tend to be significantly closer
to -1 than the calculated values.
Except at the low experimental values of $\alpha(Z)$ where the scatter
is significant, the experimental values for $\beta(N)/\alpha(Z)$ are about
$20\%$ lower in the absolute value than the model values as indicated by the ratio of these two
quantities also plotted in the middle part of the figure.
In order to see the
system dependence of $\alpha(Z)$ and $\beta(N)$ values, these values are
plotted for separate groups of fissility values in Fig.2.
The fissility is defined as $X=\frac{Z_s^2}{A_s}$, where
$Z_{s}$ and $A_{s}$ are the charges and masses of the
source which we assume to be the compound nucleus for simplicity.
We can define a combined fissility parameter between reactions (1)
and (2) as $\Sigma X=X_1+X_2$.
Larger (absolute) values of $\alpha(Z)$ and $\beta(N)$ correspond to
large values of the $\Sigma X$ parameter.
In the figure $\alpha(Z)$ and $\beta(N)$ values are separately plotted for four different
ranges of fissility group for the same data set used in Fig.1.
We see no systematic correlation with the fissility parameters in the
deviation from $\alpha(Z) = -\beta(N)$,
which might be suggestive of the fact that the Coulomb force is not
so effective for breaking the invoked invariance of the Nuclear Hamiltonian.
It would be very interesting to see if Coulomb effects become more
important for heavy colliding nuclei such as $U+U$.
One should note that a similar result is observed for
$\alpha(Z)$ and $\beta(N)$ in other IMFs, when Z and N are the same.
One can also use $\alpha$ and $\beta$ values averaged over a range of
atomic (or neutron) number, though in this case
the averaged numbers depend on the somewhat range selected.
\begin{figure}[ht]
\includegraphics[width=3.6in]
{Isoscaling_BetaVsAlpha_givenZN_4fiss_Z6Z13_fig2.eps}
\caption{\label{fig:fig_2}$\beta(N)$ vs $\alpha(Z)$ with $6 \le Z \le 13$ for different ranges of the
fissility parameter. $\Sigma X\le60.9$ (top left),
$60.9<\Sigma X\le72.2$ (top right),
$72.2<\Sigma X\le83.4$ (bottom left)
and $83.4<\Sigma X\le94.7$ (bottom right). }
\end{figure}
While the trend in Figs.\ref{fig:fig_1} and \ref{fig:fig_2} is interesting, it is important to
note that when the neutron and proton concentrations of the initial
excited source are different, two well established trends can act to
shift the balance toward symmetric matter and hence to bring the
absolute values of the observed $\alpha(Z)$ and $\beta(N)$ parameters closer
together. The first is the distillation effect in which early emission of
particles favors neutron emission over proton emission~\cite{serot,maria}.
As a result of this early emission the fragmenting system will tend
to have a higher symmetry than the initial system. The second is
secondary decay of initially excited fragments~\cite{Marie98,Hudan03}
which favors a shift toward the evaporation attractor line~\cite{Charity88}.
Thus even if the comparison of primary fragment yields
would lead to a significant difference in the two isoscaling
parameters the subsequent decay can reduce this difference.
\section *{$m$-SCALING}
Pursuing the question of phase transitions, we note that
we have previously discussed some of the present yield data within
the Landau free energy description~\cite{Bonasera08}. In this
approach the ratio of the free energy (per particle) to the temperature
is written in terms of an expansion:
\begin{equation}
\label{eq:order}
\frac{F}{T}=\frac{1}{2}a m^2+\frac{1}{4}b m^4
+\frac{1}{6}c m^6-m\frac{H}{T},
\end{equation}
where $m$ is the order parameter, $H$ is its conjugate variable and
$a-c$ are fitting parameters. In our case $m = (I/A)$.
Notice that the free energy that we have indicated
with F includes the chemical potential of neutrons and protons i.e.
$AF(m,T)=[G(N,Z)-\mu_nN-\mu_pZ]$ (compare to Eq.(\ref{eq:yield})).
We observe that the free energy is even in the exchange of
$m \rightarrow -m$ reflecting the invariance of the nuclear
forces when exchanging N and Z. This symmetry is violated by
the conjugate field $H$ which arises when the source is asymmetric
in the chemical composition. We stress that correctly m and H are
related to each other through the relation $m=-\frac{\delta F}{\delta H}$.
An immediate consequence of the application of the Landau expression of
Eq.(\ref{eq:order}) in the Modified Fisher Model is that it brings a
scaling law for m=0 isotopes. Since F(m=0,T)=0, for any T,
the yield in Eq.(\ref{eq:yield}) is given as
\begin{eqnarray} {
Y(N,Z)=Y_0A^{-\tau},
}
\label{eq:yield_m0}
\end{eqnarray}
for all reactions.
\begin{figure}[ht]
\includegraphics[width=3.6in]{021705_060606_040805_RvsA_zge0_I0_fig3.eps}
\caption{\label{fig:fig_3}Yield ratio of m=0 isotopes vs A. The yield is
normalized to that of $^{12}C$. Data from all 18 reaction
systems studied in this experiment. (top) even-even
isotopes. (middle) odd-odd isotopes are plotted.
(bottom) pairing corrected yield, Y(N,Z)/exp($\delta$), are plotted for
all m=0 isotopes. Lines in each figure are linear fitted ones.
$\tau$ values are 3.3, 2.2 and 2.8 from the top to bottom.
$a_p/T$ =2.2 is used
in the bottom.}
\end{figure}
In Fig.\ref{fig:fig_3}, yield ratios for m=0 isotopes are separately plotted as a function of A for even-even (top)
and odd-odd (middle) isotopes for all 18 reactions studied here.
In order to eliminate the effect of the constant in Eq.(\ref{eq:yield}),
which are slightly different in each reaction system, the yield is
normalized to that of the $^{12}$C in each reaction.
As seen in the figure, the yields from the different reactions are indeed
scaled well with A up to A $\sim$ 30 when even-even and odd-odd isotopes are plotted
separately. One should note that data points for a given A represent
all 18 different reactions in the figure. The slope difference between even-even
and odd-odd isotopes can be naturally attributed to the pairing effect.
However a large pairing effect is expected only at a low temperature, because it is related to the shell effect.
On the other hand the emitting sources of these isotopes are expected to be
at a high temperature. Ricciardi et al.
have given a possible explanation for this observation~\cite{Ricciardi04,Ricciardi05}.
According model simulations which they have performed the experimentally observed
pairing effect is attributed to the last chance particle decay
of the excited fragments during their cooling. This hypothesis is
also supported by our model simulations presented in a separate
paper~\cite{huang10}.
In order to take into account the pairing effect, data for
even-even and odd-odd isotopes were simultaneously fitted by the following equation,
\begin{eqnarray}
Y(N,Z)=Y_{0}A^{\tau}exp(\delta/T),
\label{eq:yield_m0_delta}
\end{eqnarray}
\begin{eqnarray}
\delta(N,Z)=\left\{\begin{array}{ll}
a_{p}/A^{1/2 }&(\textrm{odd-odd})\\
0 &(\textrm{even-odd})\\
-~a_{p}/A^{1/2 }&(\textrm{even-even} ).
\end{array}\right.
\label{eq:paringterm}
\end{eqnarray}
and the parameters $\tau$ and $a_p/T$ values was extracted. Using these extracted
parameters, the experimental yield was divided by the exponent
in Eq.(\ref{eq:yield_m0_delta}) as factor. The results are plotted in the bottom
of the figure for all isotopes with m=0. The extracted $\tau$ value is 2.8
which is larger than the normal critical exponent 2.3.
This difference may reflect either the temperature of the emitting
source is below the critical temperature or that the secondary decay processes
modify the value.
Because of the symmetries of the free energy when we take the ratio
between two different systems, $presumably$ at the same temperature
$T$ and density $\rho$, all $even$
order terms in $m$ cancel out while the $odd$ terms remain.
Those terms depend on the $external$ field $H/T$.
Taking the ratio between two systems as in Eq.(\ref{eq:yield}) we easily
obtain:
\begin{equation}
R_{12}(m)=Cexp(\Delta H/T m A),
\label{eq:mscaling}
\end{equation}
where $\Delta H/T=H_1/T-H_2/T$. We can fix the constant C by dividing
each experimental yield by the $^{12}C$ yield following
in ref.~\cite{Bonasera08}. The goal is to get $C\rightarrow1$ for
reasons that will become clear below.
Comparing the latter equation with Eq.(\ref{eq:isoscaling}) we obtain:
$\Delta H/TmA=\alpha N + \beta Z$ i.e. $\alpha=-\beta=\frac{\Delta H}{T}$.
As shown in Fig.\ref{fig:fig_1}, for the comparison for isotopes of a given Z with isotones having N equal to that Z, this relation appears to be satisfied.
The relation is valid more in general, and in fact we could
write the chemical potentials of neutrons and protons as:
\begin{equation}
\mu_nN+\mu_pZ= \mu A+H m A,
\label{eq:muu}
\end{equation}
from this relation it follows that:
\begin{equation}
2H=\mu_n-\mu_p;
2\mu=\mu_n+\mu_p.
\label{eq:muuh}
\end{equation}
All these relations show that if $m$ is an order parameter then
$\alpha=-\beta = \Delta H/T$.
\begin{figure}[ht]
\includegraphics[width=3.6in]{Isoscaling_alpha_m_Z613_fig4.eps}
\caption{\label{fig:fig_4}Experimental ratios vs m for isotopes
with $6 \le Z \le 13$
for (a) $\frac {^{64}Ni^{232}Th}{^{70}Zn^{197}Au}$,
(b)$\frac{^{64}Ni^{112}Sn}{^{64}Ni^{58}Ni}$,
(c)$\frac{^{64}Ni^{124}Sn}{^{64}Ni^{64}Ni}$,
(d)$\frac{^{64}Ni^{197}Au}{^{64}Ni^{112}Sn}$,
(e)$\frac{^{64}Ni^{124}Sn}{^{70}Zn^{58}Ni}$ and
(f)$\frac{^{64}Ni^{124}Sn}{^{70}Zn^{112}Sn}$ respectively at
$40 MeV/A$.
The lines are the results of a linear fit according to Eq.(\ref{eq:mratio}).
}
\end{figure}
The external field is given by the difference of chemical potentials
between neutrons and protons of the emitting system as expected.
>From Eq.(\ref{eq:mscaling}) we can obtain the difference between
the free energies (or alternatively the external fields) as:
\begin{equation}
\frac{-ln(R_{12}(m))}{A}=\Delta H/T m +constant.
\label{eq:mratio}
\end{equation}
Thus a plot of $-ln(R)/A$ versus m should give a linear relation
whose slope is given by $\Delta H/T$.
This linear relation is demonstrated in Figs.\ref{fig:fig_1}
and \ref{fig:fig_2} where such a plot
is obtained for different colliding systems for the isotopes in
the selected range of Z. In thatgiven range $\alpha(Z)$ increases
about 50\% on average~\cite{chen,huang10}. As discussed in references~\cite{chen,huang10},
the observed fragment Z (or N) dependence of the isoscaling parameters is mainly established
during the statistical cooling of the excited fragments.
In fact it has been demonstrated that $\alpha(Z)$ parameter extracted from
the primary fragments of the AMD simulations shows no significant
fragment Z dependence.
It should be noted that it is important to normalize the distribution
( for instance to $^{12}C$ ) as we have done in order that the
normalizing constant in front of the yield in Eq.(\ref{eq:mscaling})
is one. If not this will carry a $\frac{1}{A}$ term which might violate
the scaling.
Overall the scaling is satisfied for this set of data as seen in Fig.\ref{fig:fig_4}.
Compared to 'traditional' isoscaling where a fit is
performed for each detected charge $Z$ (or each $N$) we see that all the
data collapse into one curve.
We can further elucidate the role of the external field $H/T$ writing the Landau expansion and 'shifting' the order parameter by $m_s$ which is the position of the minimum of the free energy.
Such a position depends on the neutron to proton concentration of the source~\cite{Bonasera08}. Thus
\begin{equation}
\label{eq:shifted}
\frac{F}{T}=\frac{1}{2}a (m-m_s)^2+\frac{1}{4}b (m-m_s)^4 +\frac{1}{6}c (m-m_s)^6.
\end{equation}
Comparing to Eq.(\ref{eq:order}) we easily obtain
\begin{equation}
\label{eq:hms0}
\frac{H}{T}=(a + b m^2 + c m^4 )m_s
+(\frac{1}{2}\frac{a}{m}+\frac{3}{2}b m +\frac{5}{2}c m^3)m_s^2 + O(m_s^4)...,
\end{equation}
thus $H$ depends on the source isospin concentration though the
parameter $a, b, c$ which are terms of the free energy.
We stress that these terms refer to the free energy and $not$
to the internal symmetry energy. If b and c are of comparable magnitude to parameter a, then taking terms of a, Eq.(\ref{eq:hms0}) can be further simplified as
\begin{equation}
\label{eq:hms}
\frac{H}{T}=-a m_s +\frac{1}{2}a\frac{m_s^2}{m}+O(m_s^4)...,
\end{equation}
\section{Reconciliation of the two approaches}
Standard isoscaling results have been derived under a general grand canonical approach~\cite{Ono03,Tsang01,Botvina02}. The Landau approach should be equivalent to it under certain conditions. Experimentally the b and c values have not been established because all isotopes identified in the present data have m $<$ 0.5 except for nucleons. In the case that b and c are of comparable magnitude to parameter a, which is assumed in the derivation of eq.(\ref{eq:hms}), we easily obtain:
\begin{equation}
\label{eq:deltahms}
\frac{\Delta H}{T}mA= a \Delta m_s (N-Z) - \frac{1}{2}a (m^2_{s1}-m^2_{s2}) A =\alpha N+ \beta Z
\end{equation}
which introduces a volume term. Equating similar terms we get:
\begin{equation}
\label{eq:alphabeta}
\alpha=a \Delta m_s -\frac{1}{2}a (m^2_{s1}-m^2_{s2})
\end{equation}
where $\Delta m_s=m_{s1}-m_{s2}$. It is straightforward to demonstrate the equivalence of the last equation to eq.(\ref{eq:alpha}) derived from the grand canonical approach. In particular we get:
\begin{equation}
\label{eq:alpha+beta}
\alpha+\beta=-a (m^2_{s1}-m^2_{s2})
\end{equation}
which shows that the two approaches are equivalent and that $m$ is an order parameter if $\alpha +\beta =0$ i.e. neglecting $O(m_s^2)$ terms in the external field. In figure (\ref{fig:fig_5}) we plot $\alpha+\beta$ vs. $m^2_{s1}-m^2_{s2}$, unfortunately the error bars are rather large but we can see a systematic deviation from zero as expected from eq.(\ref{eq:alpha+beta}) for large differences in concentration.
\begin{figure}[ht]
\includegraphics[width=3.6in]{Isoscaling_Beta_Alpha_Vs_ms_werror_Z7_fig5.eps}
\caption{\label{fig:fig_5}$\alpha(Z)+\beta(N)$ vs the difference (solid circles)
in concentration for the two reaction systems for the case of Z=N=7. The dotted line is the
result of a linear fit.}
\end{figure}
This indicates that, at the level of sensitivity so far acheived with data of this type the presence of higher order terms in m is difficult to quantify. Thus, within the error bars, $m$ could be considered an order parameter when relatively neutron (or proton) rich
sources are considered. In particular, phase transitions in finite systems could be studied using the same language of macroscopic systems i.e., 'turning on and off' an external field~\cite{Bonasera08}.
\section *{ SUMMARY}
In conclusion, in this paper we have discussed scaling of ratios of
yields from different colliding systems under similar physical
control parameters, i.e. density and temperature.
A careful and precise determination of isotopic yields is needed
in order to see the features of the system near the phase transition.
There is an order parameter, m,
given by the difference in neutron and proton concentrations of the
detected fragments which leads to an expected isoscaling relation,
a direct consequence of the restored symmetry of the nuclear Hamiltonian
when exchanging neutrons with protons. The data suggest that the
Coulomb field may not significantly violate such a symmetry.
The existence of m-scaling might be a signature for near criticality
of the fragmenting system.
Other properties of the 'rich' nuclear Hamiltonian, such as pairing,
appear to result in small violations of the scaling. This is an
interesting physical aspect which deserves further and deep
investigation both theoretically
and experimentally. Also, it would be interesting to search
for m-scaling violations in heavily charged colliding systems
such as U+U. The absence of a violation in
these cases would suggest that densities and deformations of the
fragments are such that the effect of Coulomb is significantly reduced.
Studies of the other extreme case of very exotic colliding systems
would be also
be valuable to probe the effects of high 'external' field on the phase
transition. The atomic nucleus constitutes a formidable laboratory to
test our knowledge and understanding of phase transitions in a finite
system and offers a unique possibility for different quantum aspects
similar to other bosons and fermion mixtures.
A major consideration in the interpretation of the results
presented in this paper is the effect of the secondary decay
process. In the experiments excited fragments cool down to
the ground state before they are detected. The reconstruction of the
primary fragments from the experimentally observed IMFs and associated
particles is not straightforward, since multiple excited primary
fragments may be simultaneously produced in multifragmentation reactions
making the unambiguous identification of the primary fragment distribution
difficult. Indeed a major goal of the experiments from which the
present isoscaling data are taken was to employ fragment-particle
correlation measurements to reconstruct the primary fragment
distribution. The correlation data are still being analyzed~\cite{Wada05}.
\begin{acknowledgments}
This work is supported by the U.S. Department of Energy and the Robert
A. Welch Foundation under grant A0330. One of us(Z. Chen) also thanks
the \textquotedblleft100
Persons Project" of the Chinese Academy of Sciences for the support.
\end{acknowledgments}
|
\section{Introduction}
Representing collections of social relations as networks has become a
standard practice not only in sociology but also in the study of
complex systems. A social network $G = (V,E)$ is defined by the set of
$N$ nodes, $V = \{v_i\}_{i=1}^N$, which correspond to individuals, and
by the set of edges (or links) $E = \{e_{ij}=(v_i,v_j),\,v_i,v_j \in
V\}$, which correspond to some associations between the
individuals. Even though it is common to assume the associations to be
symmetric (i.e. $e_{ij} = e_{ji}$) \cite{Goodreau2007,
Onnela_PNAS_2007, Liljeros2001}, for our purposes the edges are
always directed, meaning that $e_{ij}$ is independent of $e_{ji}$. In
a \emph{weighted} network each edge is also assigned a weight. We
consider the edge weights as proxies for the strength of a
relationship, and therefore the weights are strictly positive. In our
notation $w_{ij} > 0$ is the weight of a directed edge from node $v_i$
to $v_j$, and $w_{ij} = 0$ is equivalent to saying that there is no
edge from $v_i$ to $v_j$.
We use the term \emph{reciprocity} to depict the degree of mutuality
of a relationship. A relationship with a high reciprocity is one where
both are equally interested in keeping up the relationship --- a good
example is the relationship of Romeo and Juliet in the famous play
with the same name by William Shakespeare, where the two lovers strive
to share each other's company despite their families' objections. On
the other hand, in a relationship with a low reciprocity one person is
significantly more active in maintaining the relationship than the
other. We judge reciprocity from actual communications taking place
between people.\footnote{Of course, the reciprocity deduced from
communication, even when the communication is completely known, does
not necessarily match the individuals' \emph{perceptions} of the
reciprocity.} In Shakespeare's time this would have meant counting
the number of poems passed and sonnets sung, but in our modern era it
is easier to make use of the prevalence of electronic communication.
From the sociological point of view the directionality of associations
is interesting \emph{per se}: are human relationships typically mutual
or not? If not, what possible explanations are there for the lopsided
relations? Directionality of communication might also affect for
example the spreading of mobile viruses \cite{Wang2009}, or the
spreading of information and ideas in societies. There is also
evidence that the reciprocity of a relationship is a good predictor of
its persistence \cite{HidalgoPhysA2008}.
We note that even though symmetrical associations are common in the
social networks literature, this is rarely meant as a claim of perfect
reciprocity. For many purposes the assumption of undirected edges is a
useful simplification or it follows directly from the definition of
association. This is the case for example with scientific
collaboration networks \cite{Newman2001}, where we draw a link between
two scientists if they have collaborated in writing an article. On the
other hand, some associations are intrinsically directed, like e-mail
networks \cite{Holme2005} or friendship networks. The latter was shown
explicitly in \cite{Goodreau2007} with data from the National
Longitudinal Study of Adolescent Health, where US students were asked
to name up to 5 of their best friends: only 35 \% of the roughly 7000
such nominations were reciprocated. Moreover, some undirected networks
can be considered as directed networks when examined more closely. For
example, collaboration networks could be extended by adding
information on which party initiated the collaboration; the same is
true for sexual networks \cite{Liljeros2001}.
Previous studies have mostly considered reciprocity as a global
feature of a directed, unweighted network. They are based on the
classical definition according to \cite{Wasserman1994} that quantifies
reciprocity as $r = \frac{L^{\leftrightarrow}}{L}$ where $L = \sum_{i
\neq j}a_{ij}$ is the total number of (directed) links in the
network and $L^{\leftrightarrow} = \sum_{i \neq j} a_{ij}a_{ji}$ is
the number of the mutual edges (the latter sum goes over all pairs of
nodes). Here $a_{ij}$ are the elements of the adjacency matrix such
that $a_{ij}=1$ if there is a directed link from node $v_i$ to $v_j$,
and $a_{ij} = 0$ otherwise. Defined this way, $r = 0$ if there are no
bidirectional links, and $r=1$ when all links in the network are
bidirectional. For example \cite{Newman2002} finds that one network
constructed from email address books has $r = 0.231$.
One problem with this definition of reciprocity is that it depends on
the density of the network. A large value of $r$ is more significant
when encountered in a sparse network, while a small $r$ is a
surprising finding in a dense network. The value of $r$ must thus be
compared to the network density $\bar{a} = \frac{L}{N(N-1)}$, which also
equals the reciprocity of a fully random
network. In \cite{Garlaschelli2004} this is taken further by combining $r$
and $\bar{a}$ into a single measure of reciprocity, $\rho =
\frac{r-\bar{a}}{1-\bar{a}}$. This idea is extended in
\cite{Zamora-Lopez2008} by comparing $r$ in the original network to
that in randomized networks when the degree sequence or various degree
correlations are held constant. It turns out that in most networks
this is enough to explain most of the reciprocity.
The unweighted treatment of reciprocity however has a severe problem
when dealing with social networks: a truly unidirectional social
contact is an extremely rare thing to find in everyday social
environment. When such edges do occur, for example in the
above-mentioned study where the subjects were asked to name a fixed
number of acquaintances \cite{Goodreau2007}, the unidirectional links
are mostly artifacts of the research method. They can however be
interpreted as a sign of an underlying edge bias, a cue that the
relationship is not seen alike by the two people involved.
Unlike in the unweighted case, edge weights allow us to study
reciprocity as a property of a single edge instead of the full
network. To this end we define the \emph{edge bias} between nodes
$v_i$ and $v_j$ as
\[
b_{ij} = \frac{w_{ij}}{w_{ij}+w_{ji}}.
\]
The edge bias is the simplest possible measure for studying
reciprocity. Note that because $b_{ij} + b_{ji} = 1$, the distribution
of $b_{ij}$ for all edges in the network is symmetric around $0.5$ and
therefore it suffices to study only the range $b_{ij} \in [0.5,\,1]$.
\section{Data}
We study the reciprocity of communication in a mobile phone data set
consisting of 350 million calls between 5.3 million customers made
during a period of 18 weeks. Mobile phone calls provide an excellent
proxy for studying this kind of reciprocity because phone calls are at
the same time directed and undirected: The calls are directed because
it is the caller who decides to make the call and invest his time (and
often money) in that particular relationship, but since the
conversation during the call is undirected, there is no immediate
reason for the recipient to call back shortly after. Consider the
difference to SMS messages or emails where a conversation means
sending a sequence of reciprocated messages, making it more difficult
to study reciprocity by simply counting the number of messages.
To remove possibly spurious contacts, i.e. those that are more likely
to be chance encounters instead of true social relationships, we have
preprocessed the data by removing all unidirectional edges. Thus the
unweighted reciprocity according to the definition above would be
$r=1$, and therefore the methods for analyzing the reciprocity of
unweighted, directed networks are not applicable.\footnote{To be
exact, $r \approx 1$ because any communication, either SMS message
or call, was sufficient to make an edge bidirectional, and thus
there are some edges where only one person has made calls.} The
weight $w_{ij}$ is the total number of calls customer $v_i$ made to
$v_j$ during the whole 18 week period.
\begin{figure} \centering
\includegraphics[width=7cm]{figures/total_degree_dist_RecArt}
\caption{Total degree distribution of the mobile phone data set
separately for prepaid and postpaid users. Inset: The distribution
until $k=30$ in semi-logarithmic coordinates.}
\label{fig:degree_distribution}
\end{figure}
The customers can be split into two groups based on how they pay for
their mobile phone service: \emph{prepaid} users pay for their usage
beforehand, \emph{postpaid} users pay afterword. While this initial
difference might seem small, it results in major differences in the
basic statistics and it is therefore necessary to analyze the two user
groups separately. The statistics of calls between prepaid and
postpaid users would mostly just reflect the differences in degree and
strength distributions (discussed below). Thus in this article we only
study calls between pairs of prepaid users and calls between pairs of
postpaid users. The two user groups are approximately equal in size.
The basic statistics of the network often imprint also more complex
statistics, and it is therefore a good idea to first go through some
basic distributions. The \emph{degree} $k_i$ of a node is defined as
the number of neighboring nodes. Because all edges in our network are
bidirectional after preprocessing, the out-degree (number of people
called) always equals the in-degree (number of people from whom a call
has been received). The degree distributions of the network, shown in
Figure \ref{fig:degree_distribution} separately for prepaid and
postpaid users, have long power-law tails. The average degree was
found to be 3.41 for prepaid users and 5.15 for postpaid users.
The \emph{strength} of a node is defined as the sum of weights of
adjacent edges. For a directed network we can define both the
out-strength $s_i^{\textrm{out}} = \sum_j w_{ij}$, which is the total
number of calls made, and the in-strength $s_i^{\textrm{in}} = \sum_j
w_{ji}$, the total number of calls received. The strength
distributions in Fig. \ref{fig:strength_distribution} turned out also
to be fat-tailed, and we can see that postpaid users make much
more calls. In fact, postpaid users make on average 10 times more
calls than prepaid users. We can also see that prepaid users receive
more calls than they make, while the most active postpaid users make
more calls than they receive.
\begin{figure} \centering
\subfigure[]{
\includegraphics[width=7cm]{figures/strength_dist_RecArt}
\label{fig:strength_distribution}
}
\subfigure[]{
\includegraphics[width=7cm]{figures/weight_distributions_RecArt}
\label{fig:weight_distribution}
}
\caption{\textbf{(a)} The in- and out-strength distributions for the two users
types. \textbf{(b)} The total weight distribution for edges between
the two user types.}
\end{figure}
Figure \ref{fig:weight_distribution} shows the total edge weight
distribution for edges between users of the same or different
type. There are more calls between users of the same type, which
suggests that the user type is correlated between acquaintances.
The data set only includes the customers of a single mobile phone
operator that has about 20 \% market share in the country in
question. While there are calls from and to people outside this
customer base, this has no skewing effect on our study because our
analysis concentrates on communication between the known customers,
and this information is completely known.
\section{Results}
\label{sec:results}
Figures \ref{fig:bias_dist_prepaid} and \ref{fig:bias_dist_postpaid}
show the distribution of $b_{ij}$ for prepaid and postpaid users as a
function of total edge weight. We can see that even though values
around 0.5 are more common, there are still many edges with larger
values of $b_{ij}$. This becomes more obvious in Figure
\ref{fig:cumulative_bias_by_weight} that shows the cumulative
distribution of edge bias for different weight ranges. For example
among prepaid users the edges with $b_{ij} \geq 0.8$, i.e. edges where
one participant makes more than 80 \% of all calls, make up over 25 \%
of all the edges even when the edge has over 100 calls. For postpaid
users the numbers are somewhat smaller.
Naturally, we should not expect to find \emph{perfect} reciprocity on
all edges, but to justify treating social networks as undirected there
should be some tendency towards $b_{ij}=0.5$. But how much? The
simplest step away from perfectly reciprocal relationships is to
assume each dyad to be reciprocal in the probabilistic sense: both
participants are equally likely to make a call. Thus, even though
edges with only few calls could still be very biased, edges with 50 or
more calls should have $b_{ij}$ very close to $0.5$. This idea is also
supported by Figure \ref{fig:cumulative_bias_by_weight}, where the
distributions of high weight edges are somewhat more concentrated
around $0.5$.
\begin{figure} \centering
\subfigure[Prepaid]{
\includegraphics[width=7cm]{figures/bias_stats_call_count_pre2pre}
\label{fig:bias_dist_prepaid}
} \subfigure[Postpaid]{
\includegraphics[width=7cm]{figures/bias_stats_call_count_post2post}
\label{fig:bias_dist_postpaid}
}
\caption{The distribution of edge bias values as a function of edge
weight for \textbf{(a)} prepaid and \textbf{(b)} postpaid
users. Each vertical strip in the figure is a distribution and sums
up to one. When $w_{ij} + w_{ji}$ is small, we can see that the edge
bias is strongly quantized; for example if the total weight is 5,
there are only three possible values for the edge bias in the range
$b_{ij} \in [0.5,\,1]$: $\frac{3}{5} = 0.6$, $\frac{4}{5} = 0.8$ or
$\frac{5}{5} = 1.0$. The white area on right on the other hand is
caused by missing values; there are not many edges with over 2000
calls in 18 weeks.}
\end{figure}
The hypothesis about equal likelihood for making a call implies that
the edge bias values should follow a binomial distribution with
$p=0.5$ and $n=w_{ij}+w_{ji}$. However, Figure
\ref{fig:binomial_comparison} shows that such a binomial distribution
is far from truth: if the calls were evenly distributed between both
participants, there should be almost no edges with $b_{ij} > 0.7$.
But is it reasonable to expect the edge bias values to be binomially
distributed in the first place? As shown in
Fig. \ref{fig:strength_distribution}, the node strengths vary widely,
which itself might be enough to force large deviations from $b_{ij} =
0.5$ --- in other words the relationships would only appear biased
because some people tend to call more than others. To test this claim
we redistribute the total out-strength of each node onto its already
existing edges so as to make the edge bias values as close to 0.5 as
possible (see Appendix for details). It turns out that it is in fact
possible to significantly reduce the number of high edge bias values,
especially among postpaid users, as shown in
Fig. \ref{fig:reference_comparison}. This proves that the strength
distribution is not a sufficient explanation for the large edge bias
values.
\begin{figure} \centering
\subfigure[]{
\includegraphics[width=7cm]{figures/cumulative_call_count_bias_by_weight_RecArt}
\label{fig:cumulative_bias_by_weight}
}
\subfigure[]{
\includegraphics[width=7cm]{figures/cumulative_call_count_bias_dist_binComp}
\label{fig:binomial_comparison}
}
\subfigure[]{
\includegraphics[width=7cm]{figures/cumulative_call_count_bias_by_overlap_RecArt}
\label{fig:cumulative_bias_by_overlap}
}
\subfigure[]{
\includegraphics[width=7cm]{figures/cumulative_call_count_bias_dist_refComp}
\label{fig:reference_comparison}
}
\caption{\textbf{(a)} The cumulative edge bias distribution
separately for different ranges of total edge weight. The solid
lines are for prepaid users and the dotted lines for postpaid
users, lighter color means higher edge weight. \textbf{(b)} The
cumulative edge bias distribution for prepaid and postpaid users
compared to the cumulative binomial distribution. In theory the
binomial reference is slightly different for prepaid and postpaid
users due to different edge weight distributions, but in practice
the difference is negligible. \textbf{(c)} The cumulative edge
bias distribution separately for different ranges of overlap. The
solid lines are for prepaid users and the dotted lines for
postpaid users, lighter color means higher overlap. It would seem
that larger overlap implies edge bias values more concentrated
around 0.5, but as shown in Figure
\ref{fig:bias_by_weight_and_overlap}, this is a spurious
correlation. \textbf{(d)} The cumulative edge bias distribution
for prepaid and postpaid users compared to the references where
the out-strength of each node has been redistributed in an attempt
to equalize edges. In all plots only edges with $w_{ij} + w_{ji}
\geq 20$ are included to avoid excessive quantization of the
cumulative distribution. With 18 weeks of data this requirement
translates to an average of roughly one call (in either direction)
per week. These edges contain about 85 \% of all calls in the
network.}
\end{figure}
Since it appears that the edge bias cannot be explained away by
properties of the node but is instead an intrinsic property of the
edge itself, it is interesting to see whether it correlates with other
local features of the network. Is there for example a connection
between the edge bias and the community structure of the network? We
measure local community structure with \emph{edge overlap}, defined as
$O_{ij} = n_{ij}/((k_i- 1) + (k_j - 1) - n_{ij})$, where $n_{ij}$ is
the number of common neighbors of nodes $v_i$ and $v_j$. Thus the edge
overlap gives the fraction of common neighbors out of all possible
common neighbors, and is therefore larger in denser parts of the
network.
Looking at Figure \ref{fig:cumulative_bias_by_overlap} it seems that
the edge bias does indeed correlate with overlap: larger overlap
implies an edge bias more concentrated around 0.5 for both prepaid and
postpaid users. However, Figure \ref{fig:bias_by_weight_and_overlap}
shows that the observed effect is in fact a spurious result of two
other correlations. The first correlation is the one between edge bias
and weight, shown in Fig. \ref{fig:cumulative_bias_by_weight}. The
second correlation is the \emph{Granovetter hypothesis}
\cite{GranovetterWeakTies1973}, which states that edges between
communities should be weaker than those inside communities. It was
shown in \cite{Onnela_PNAS_2007} that in a mobile phone network this
is indeed the case: edge weights are higher in denser parts of the
network.
\begin{figure} \centering
\includegraphics[width=7cm]{figures/weight_overlap_bias_call_count_post2post}
\caption{The average edge bias as a function of the edge weight and
overlap for postpaid users (the corresponding plot for prepaid users
is qualitatively similar). The color tells the average edge bias for
edges with given total weight and edge overlap. The average edge
bias changes with the edge weight (moving left to right) but does
not significantly change with the edge overlap (moving down to up)
given the edge weight.}
\label{fig:bias_by_weight_and_overlap}
\end{figure}
Finally, Figures \ref{fig:degdeg_bias_call_count_prepaid} and
\ref{fig:degdeg_bias_call_count_postpaid} show that there is a
connection between the edge bias and the degrees of the caller and the
recipient. This simply confirms the everyday observation that people
with many contacts tend to be active in keeping up those contacts, as
opposed to simply waiting for the others to call them. While for prepaid
users this observation results at least partly from the fact that
average out-weight increases with degree
(Fig. \ref{fig:avg_out_weight_prepaid}), this is not the case with
postpaid users. In fact, of the edges between two postpaid users the one
with a larger degree is on average more active, even though the
average edge weight decreases with degree
(Fig. \ref{fig:avg_out_weight_postpaid}).
\begin{figure} \centering
\subfigure[Prepaid]{
\includegraphics[width=7cm]{figures/degdeg_bias_call_count_weighted_prepaid}
\label{fig:degdeg_bias_call_count_prepaid}
}
\subfigure[Prepaid]{
\includegraphics[width=7cm]{figures/avg_weight_call_count_prepaid_median_RecArt}
\label{fig:avg_out_weight_prepaid}
} \\
\subfigure[Postpaid]{
\includegraphics[width=7cm]{figures/degdeg_bias_call_count_weighted_postpaid}
\label{fig:degdeg_bias_call_count_postpaid}
}
\subfigure[Postpaid]{
\includegraphics[width=7cm]{figures/avg_weight_call_count_postpaid_median_RecArt}
\label{fig:avg_out_weight_postpaid}
}
\caption{\textbf{(a,c)} The (weighted) average edge bias as function
of caller degree $k_c$ and recipient degree $k_r$ for (a) prepaid
and (c) postpaid users. The color tells the weighted average of the
edge bias values have been weighted by total edge weight. The
weighted average puts more emphasis on more active edges and is
therefore a better indicator for the mass of calls taking place in
the network. \textbf{(b,d)} The average out-weight as a function of
node degree for (b) prepaid and (d) postpaid users. The solid line
is the median, dashed lines give the 25th and 75th and the dotted
lines the 5th and 95th percentiles. One can see from (a) and (c)
that on average the greater degree participant is more active. While
this effect is much more subtle for postpaid users, it is partly
because it goes against the average edge weight shown in (d).}
\end{figure}
\section{Discussion}
We have shown that highly biased edges where one participant is much
more active than the other are abundant in a mobile phone
network. Although the strength distribution can explain some of the
observed high bias values, it is by no means a sufficient explanation.
Making a phone call is always an investment in a relationship: not
only does it cost time and money to the caller, but the caller must
also make a conscious decision to make the call. While people are in
general known to strive to make their relationships reciprocal, we can
observe plenty of relationships with a striking lack of reciprocity.
It was mentioned earlier that the fact that the data only includes 20
\% of the population has no adverse effect on the data. This should
now be quite obvious, since we have only looked at the edges of which
we have full information. However, we can state something even
stronger: it is likely that any additional data would only
\emph{improve} the reference calculated above, because there would be
more edges to choose from during the optimization. This of course
assumes that the edge bias on the missing edges is distributed
similarly as in the current data, which we feel is a very reasonable
assumption.
We have so far deferred the discussion about whether the results can
be extended to human relations in general. There are obviously many
factors that influence the reciprocity of a relationship, as is
evident from the differences between prepaid and postpaid users. We
note that these difference do not necessarily stem straight from the
differences in paying. For example, since the prepaid service is used
more often by young people, our findings suggest that the relations
among young people are more biased. Obviously hypotheses like this one
need more research until we can say anything conclusive.
Data sets including communication through several different channels,
such as email, online chatting and real life conversations, would be
especially well suited for studying reciprocity, but unfortunately
such data sets are difficult to acquire in the large scale
required. The methods used here can however be readily applied to any
such data, as long as the data includes information on who initiated
the communication.
\section*{Acknowledgment}
This work is partially supported by the Academy of Finland, the
Finnish Center of Excellence programme 2006--2011, proj. 129670. We
would also like to thank prof. Albert-L\'{a}szl\'{o} Barab\'{a}si of
Northeastern University for providing us access to the unique mobile
phone data set and continuing collaboration thereof.
|
\section{Introduction}
Quantum field theories defined in a noncommutative space have been under intense scrutiny in the last years \cite{Dou,Sza}. The
outcome of these investigations have unveiled various unusual and intriguing aspects which are consequences of their inherent nonlocality. Among these properties, the most peculiar one is the transmutation of part of the ultraviolet divergences into infrared ones, a property that has been called infrared/ultraviolet mixing \cite{MinRaa}. From a technical viewpoint, the mixing is due to the separation of the contributions of Feynman diagrams in parts nonplanar, which are ultraviolet finite but may present an infrared singularity, and planar, which may have only ultraviolet divergences. Aside the potentially dangerous character of the infrared divergences, the mere separation of the amplitudes in planar and nonplanar parts may obstruct the ultraviolet renormalization of noncommutative theories.
In the commutative setting, it is well known that Slavnov-Taylor (ST) identities \cite{slavnov} play a fundamental role in the renormalization of non-Abelian gauge theories \cite{slavnov,marciano}. It is therefore essential to verify to what extension these identities are affected by the noncommutativity of the underlying space. In this work we will present a detailed analysis of the one-loop ST identities in noncommutative QED$_4$. As we will explicitly verify, there are no anomalies and the usual renormalization procedure is not basically modified.
We would like to point out some relevant studies on the subject. For the pure noncommutative $U(N)$ Yang-Mills model, the compatibility of dimensional renormalization with the ST identities have been verified in \cite{Martin} up to one-loop order. Reference \cite{Jabbari} contains an explicit on-shell verification of the one-loop ST identity for the trilinear fermion-photon vertex. In the tree
approximation, the identities have been verified in various scattering processes in \cite{Mariz}. They were also used in \cite{Frenkel} to investigate the dependence of the two point function of the gauge field on the gauge parameter. To prove the absence of radiative corrections to the Chern-Simons coefficient, the axial gauge identities were used and explicitly verified in a one-loop calculation \cite{Das}.
This work is organized as follows. In section II we introduce our basic notation and the Feynman rules for noncommutative QED$_4$. Section III provides a formal derivation for the ST identities. In particular, using these relations the longitudinal part of photon propagator is fixed and the identities for the vectorial fermion-photon and triple photon vertex functions are presented. In Section IV these identities are subjected to a detailed analysis taking in consideration
the counterterms needed to control the ultraviolet behavior. Section V contains some final comments and a discussion of our results.
\section{Noncommutative QED$_4$}
Classically, the noncommutative QED$_4$ is described by the action
\begin{equation}\label{QED}
S_{INV} = \int d^4x \left[ - \frac14 F_{\mu\nu} \star F^{\mu\nu} + \bar\psi \star (i \slashed{D} - m)\psi \right],
\end{equation}
where $F_{\mu\nu} = \partial_\mu A_\nu - \partial_\nu A_\mu - ie [A_\mu,A_\nu]_\star$, with $[A_\mu,A_\nu]_\star=A_{\mu}\star
A_{\nu}- A_{\nu}\star A_{\mu}$, is the field strength,
$D_\mu \psi = \partial \psi - ie A_\mu \star \psi$ is a gauge covariant derivative and the star (Moyal) product is defined by
\begin{equation}
\phi_{1}(x)\star\phi_{2}(x)\equiv e^{\frac{i}{2}\xi\theta^{\mu
\nu}\frac{\partial\;\;}{\partial x^\mu}\frac{\partial\;\;}{\partial y^\mu}}\phi_{1}(x)\phi_{2}(y)|_{y=x},
\end{equation}
where $\theta_{\mu\nu}$ is a real antisymmetric matrix and $\xi$ is a parameter which sets the strength of the noncommutativity.
The above action is invariant under the gauge transformations
\begin{eqnarray}
\delta A_\mu &=&\frac1e D_{\mu} \Lambda\equiv\frac1e (\partial_\mu \Lambda - ie[A_\mu,\Lambda]_\star), \nonumber\\
\delta \psi &=& i \Lambda \star \psi \quad(\delta \bar\psi = -i \bar\psi \star \Lambda).
\end{eqnarray}
To complete the quantum version of the model, we need to add to (\ref{QED}) a gauge fixing, $S_{GF}$, and the corresponding Faddeev-Popov, $S_{FP}$, actions. For the general class of Lorentz gauges in which we will work
\begin{equation}
S_{GF} + S_{FP} = \int d^4 x \left[ - \frac1{2\alpha} (\partial_\mu A^\mu)_\star^2 + \partial_\mu \bar C \star (\partial^\mu C - ie [A^\mu,C]_\star) \right],
\end{equation}
where $\alpha$ is the gauge fixing parameter. As it happens in the commutative gauge theories, the total action $S=S_{INV}+S_{GF}+S_{FP}$ is not invariant under gauge transformations anymore but instead
has a BRST symmetry such that
\begin{eqnarray}
\delta A_\mu &=& -\frac1e (\partial_\mu C - ie[A_\mu,C]_\star) \lambda, \nonumber\\
\delta \psi &=& - i C\lambda \star \psi \quad(\delta \bar\psi = i \bar\psi \star C\lambda),\nonumber\\
\delta C &=& i C \star C\lambda, \nonumber\\
\delta \bar C &=& - \frac{1}{\alpha e} (\partial_\mu A^\mu)\lambda,\label{BRS1}
\end{eqnarray}
where $\lambda$ is a constant Grassmmanian parameter. At a formal level, the invariance of the action under these
transformations imply in relations between the Green functions as it will be shortly verified. For an explicit
calculation, we will need the Feynman rules for the model which are fixed as follows.
First, the free propagators are the same as in the commutative version of the model, i.e.,
\begin{eqnarray}
\raisebox{-0.4cm}{\incps{fermion.eps}{-1.5cm}{-.5cm}{1.5cm}{.5cm}}
&=& \frac{i}{\slashed{p}-m}, \\
\raisebox{-0.4cm}{\incps{photon.eps}{-1.5cm}{-.5cm}{1.5cm}{.5cm}} &=& -
\frac{i}{p^2} \left[ g^{\mu\nu} - (1-\alpha) \frac{p_\mu p_\nu}{p^2} \right], \;\mathrm{and} \\
\raisebox{-0.4cm}{\incps{ghost.eps}{-1.5cm}{-.5cm}{1.5cm}{.5cm}}
&=& \frac{i}{p^2},
\end{eqnarray}
for the fermion, photon and ghost field propagators, respectively. Introducing the notation
$p\wedge k \equiv \frac12\xi\theta^{\mu\nu}p_\mu k_\nu$, we determine the
vertices as being
\begin{eqnarray}
\raisebox{-0.4cm}{\incps{vertex.eps}{-1.5cm}{-0.5cm}{1.5cm}{1.5cm}}
&=& -ie \gamma^\mu e^{i p\wedge k}, \nonumber\\ \\
\raisebox{-0.4cm}{\incps{vertexp.eps}{-1.5cm}{-0.5cm}{1.5cm}{1.5cm}}
&=& 2e \sin(p\wedge q)\gamma^{\mu\nu\alpha}(p,q,k), \nonumber\\ \\
\raisebox{-0.4cm}{\incps{vertexpp.eps}{-1.5cm}{-0.5cm}{1.5cm}{1.5cm}}
&=& -4ie^2 \left[(g^{\mu\beta}g^{\alpha\nu} - g^{\mu\nu}g^{\alpha\beta})\sin(p \wedge r)\sin(q \wedge k) \right.\nonumber\\
&& + (g^{\mu\nu}g^{\alpha\beta} - g^{\mu\alpha}g^{\nu\beta})\sin(q \wedge p)\sin(r \wedge k) \\
&& + \left.(g^{\mu\alpha}g^{\nu\beta} - g^{\mu\beta}g^{\alpha\nu})\sin(p \wedge k)\sin(r \wedge q)\right], \nonumber\\
\raisebox{-0.4cm}{\incps{vertexg.eps}{-1.5cm}{-0.5cm}{1.5cm}{1.5cm}}
&=& 2e k^\mu \sin(p \wedge k), \nonumber\\
\end{eqnarray}
where $\gamma^{\mu\nu\alpha}(p,q,k)=(p-q)^\alpha g^{\mu\nu}+(q-k)^\mu g^{\nu\alpha}+(k-p)^\nu g^{\alpha\mu}$.
\section{Slavnov-Taylor identities for the generating functionals: formal aspects}
Following the standard procedure adopted in commutative gauge theories, we start by considering the generating
functional for the Green functions of the basic fields and their BRST variations,
\begin{equation}
Z[J,\eta,\bar\eta,\zeta,\bar\zeta;K,v,\omega,\bar\omega] = \int DA_\mu D\psi D\bar\psi DC D\bar C e^{i(S +S_{source})} \label{Z},
\end{equation}
where $S$ was given in the previous section and
\begin{eqnarray}
S_{source} &=& \int d^4 x \left [J_\mu \star A^\mu + \bar\eta \star \psi + \bar\psi \star \eta + \bar\zeta \star C + \bar C \star \zeta \right. \nonumber\\
&&\left.+ K_\mu \star \frac1e (\partial^\mu C - ie [A^\mu,C]_\star) + iv \star C \star C + i\bar\omega \star C \star \psi + i\bar\psi \star C \star \omega\right].
\end{eqnarray}
The invariance of the functional integral (\ref{Z}) under the field-coordinate transformation (\ref{BRS1}) and the nilpotency of that variations imply the ST identity
\begin{equation}\label{gfc}
\int d^4x \left(J_\mu \star \frac{\delta W}{\delta K_\mu} - \bar\zeta \star \frac{\delta W}{\delta v} - \frac{1}{\alpha e}\partial_\mu \frac{\delta W}{\delta J_\mu}\star\zeta -\bar\eta \star \frac{\delta W}{\delta\bar\omega} + \frac{\delta W}{\delta\omega} \star \eta\right) = 0,
\end{equation}
where $W=-i \ln \, Z$ is the generating functional for the connected Green functions. Furthermore, by subjecting the
functional integral to an arbitrary variable change $\delta \bar C$, we may derive that
\begin{equation}\label{xi}
\zeta = e \partial_\mu \frac{\delta W}{\delta K_\mu}.
\end{equation}
As usual, the generating functional $\Gamma$ of proper (one-particle-irreducible) vertex functions is obtained by a
Legendre transformation
\begin{eqnarray}
W[J,\eta,\bar\eta,\zeta,\bar\zeta;K,v,\omega,\bar\omega] &=& \Gamma[A_{cl},\psi_{cl},\bar\psi_{cl},C_{cl},\bar C_{cl};K,v,\omega,\bar\omega] \nonumber\\
&&+ \int d^4x \left( J_\mu \star A^{\mu}_{cl} + \bar\eta \star \psi_{cl} + \bar\psi_{cl} \star \eta + \bar\zeta \star C_{cl} + \bar C_{cl} \star \zeta \right),
\end{eqnarray}
where we have introduced the classical fields
\begin{equation}
A^{\mu}_{cl}= \frac{\delta W}{\delta J_\mu},\qquad \psi_{cl} = \frac{\delta W}{\delta \bar \eta},\qquad \bar \psi_{cl} =- \frac{\delta W}{\delta \eta},\qquad C_{cl}=\frac{\delta W}{\delta \bar \zeta}, \qquad \bar C_{cl}=-\frac{\delta W}{\delta \zeta}.
\end{equation}
From these definitions, it follows that
\begin{equation}
\frac{\delta\Gamma}{\delta A_{cl\,\mu}} = -J^\mu,\;\; \frac{\delta\Gamma}{\delta\psi_{cl}} = \bar\eta,\;\; \frac{\delta\Gamma}{\delta\bar\psi_{cl}} = -\eta,\;\; \frac{\delta\Gamma}{\delta C_{cl}} = \bar\zeta,\;\; \frac{\delta\Gamma}{\delta\bar C_{cl}} = -\zeta.
\end{equation}
In terms of $\Gamma$ the identities (\ref{gfc}) and (\ref{xi}) become
\begin{equation}
\int d^4x \left(\frac{\delta\Gamma}{\delta A^{\mu}_{cl}} \star \frac{\delta\Gamma}{\delta K_\mu} + \frac{\delta\Gamma}{\delta C_{cl}} \star \frac{\delta\Gamma}{\delta v} - \frac{1}{\alpha e}(\partial_\mu A^{\mu}_{cl}) \star \frac{\delta\Gamma}{\delta\bar C_{cl}} + \frac{\delta\Gamma}{\delta\psi_{cl}} \star \frac{\delta\Gamma}{\delta\bar\omega} + \frac{\delta\Gamma}{\delta\omega} \star \frac{\delta\Gamma}{\delta\bar\psi_{cl}}\right) = 0\label{gama}
\end{equation}
and
\begin{equation}
i\frac{\delta\Gamma}{\delta\bar C_{cl}} = - e \partial_\mu \frac{\delta\Gamma}{\delta K_\mu}.
\end{equation}
The identity (\ref{gama}) can be simplified by redefining $\Gamma$:
\begin{equation}
\Gamma\rightarrow \Gamma- \frac{1}{2\alpha}\int d^4x (\partial_\mu A^{\mu}_{cl})^2
\end{equation}
so that we obtain
\begin{equation}
\int d^4x \left[\frac{\delta\Gamma}{\delta A^{\mu}_{cl}} \star
\frac{\delta\Gamma}{\delta K_\mu} + \frac{\delta\Gamma}{\delta C_{cl}}
\star \frac{\delta\Gamma}{\delta v} + \frac{\delta\Gamma}{\delta\psi}
\star \frac{\delta\Gamma}{\delta\bar\omega} +
\frac{\delta\Gamma}{\delta\omega}
\star \frac{\delta\Gamma}{\delta\bar\psi}\right] = 0
\end{equation}
and
\begin{equation}
i\frac{\delta\Gamma}{\delta\bar C_{cl}} + e \partial_\mu \frac{\delta\Gamma}{\delta K_\mu} = 0.
\end{equation}
Let us now consider some specific applications of the above identities.
\subsection{The photon propagator}
As a first application of the identities derived in the previous section, we will now prove that the longitudinal
part of the photon propagator is not modified by radiative corrections. To this end, we twice
differentiate the generating functional of the connected Green functions (\ref{gfc}) with respect to $\zeta(y)$ and $J^\nu(z)$ and set all sources equal to zero, which gives
\begin{equation}
-\frac1{\alpha e} \partial_y^\mu \frac{\delta^2 W}{\delta J^\nu(z)J^\mu(y)}\left | + \frac{\delta^2 W}{\delta\zeta(y)\delta K^\nu(z)}\right |=0,
\end{equation}
\noindent
where we have introduced the notation ${\cal O}\vert $ to imply that the object ${\cal O}$ at the left of the vertical bar
has to be calculated with all sources equal to zero.
But, from Eq. (\ref{xi}) it follows that
\begin{equation}
\partial_z^\nu \frac{\delta^2 W}{\delta\zeta(y)\delta K^\nu(z)}\left | = \frac{1}e \delta(z-y)\right.
\end{equation}
so that the photon propagator $D_{\mu\nu}(z-y)=-i \frac{\delta^2 W}{\delta J^\nu(z)\delta J^\mu(y)}\left |\right.$ must satisfy
\begin{equation}
\partial_z^\mu \partial_y^\nu D_{\mu\nu}(z-y) =- i\alpha \delta(z-y),
\end{equation}
which in momentum space becomes
\begin{equation}\label{WTf}
q^\mu q^\nu D_{\mu\nu}(q) =- i\alpha.
\end{equation}
\noindent
Now, compatibility with this constraint requires the propagator to have the general form
\begin{equation}
D_{\mu\nu}(q)=\left( g^{\mu\nu} - \frac{q^\mu q^\nu}{q^2} \right) D_\mathrm{T}(q^2) + \frac{\tilde q^\mu\tilde q^\nu}{\tilde q^2} D_\theta(q^2)-\frac{i\alpha}{q^2}\frac{q^\mu q^\nu}{q^2}.\label{1}
\end{equation}
Notice that, because of the charge conjugation properties \cite{Frenkel,Jabbari1}, terms of the type
\begin{equation}
\frac{\tilde q^\mu\ q^\nu+\tilde q^\nu\ q^\mu }{\tilde q^2}
\end{equation}
\noindent
are not allowed in the decomposition (\ref{1}). Thus the longitudinal part of the propagator is the same as in the free approximation. Notice also that
\begin{equation}\label{foton}
q^\mu D_{\mu\nu}(q) = - i\frac{\alpha q_\nu}{q^2},
\end{equation}
which will be useful in the next section when we will analyze the ST identity for the vectorial vertex function.
\subsection{The vectorial vertex function}
The ST identity for the vectorial fermion-photon vertex, the proper part of $\langle 0|T(\psi\bar\psi A_\mu)|0\rangle$, can be derived by turning off all the sources after differentiating the functional equation (\ref{gfc}) with respect the sources $\eta(y)$, $\bar\eta(x)$, and $\zeta(z)$. The result is
\begin{equation}
\frac1{\alpha e} \partial_z^\mu \frac{\delta^3 W}{\delta\bar\eta(x)\delta\eta(y)\delta J^\mu(z)}\left | = \frac{\delta^3 W}{\delta\zeta(z)\delta\bar\eta(x)\delta\omega(y)}\right | - \left.\frac{\delta^3 W}{\delta\zeta(z)\delta\eta(y)\delta\bar\omega(x)}\right | ,
\end{equation}
or, equivalently,
\begin{eqnarray}
\frac1{\alpha e} \partial_z^\mu \langle 0|T(\psi(x)\bar\psi(y)A_\mu(z))|0\rangle &=& i \langle 0|T(\bar C(z)\psi(x)\bar\psi(y)\star C(y))|0\rangle \nonumber\\
&&- i \langle 0|T(\bar C(z)\bar\psi(y) C(x)\star\psi(x))|0\rangle,
\end{eqnarray}
i.e.,
\begin{eqnarray}\label{vertex}
\frac1{\alpha e} \partial_z^\mu \langle 0|T(\psi(x)\bar\psi(y)A_\mu(z))|0\rangle &=& i e^{i \partial_x\wedge\partial_{\hat x}} \langle 0|T(\psi(\hat x)\bar\psi(y)C(x)\bar C(z))|0\rangle|_{\hat x=x} \nonumber\\
&& - i e^{i \partial_y \wedge\partial_{\hat y}}\langle 0|T(\psi(x)\bar\psi(y) C(\hat y)\bar C(z))|0\rangle|_{\hat y=y},
\end{eqnarray}
where $\partial_x\wedge\partial_{\hat x}=\frac12\xi\theta^{\mu\nu}\frac{\partial\;\;}{\partial x^\mu}\frac{\partial\;\;}{\partial \hat x^\nu}$. Notice that as consequence of this identity $$\delta \langle 0|T(\psi(x)\bar\psi(y)\bar C(z))|0\rangle = 0.$$ We may translate the above equations into identities for the proper, one-particle irreducible, vertex functions. These functions are given by
\begin{eqnarray}
&& \langle 0|T(\psi(x)\bar\psi(y)A_\mu(z))|0\rangle \nonumber\\
&&\quad\quad\quad\quad = - \int d^4x'd^4y'd^4z' S_F(x-x')\Gamma^\nu(x',y',z')S_F(y'-y)D_{\mu\nu}(z-z'), \\
&& i e^{i \partial_x \wedge\partial_{\hat x}} \langle 0|T(\psi(\hat x)\bar\psi(y)C(x)\bar C(z))|0\rangle|_{\hat x=x} \nonumber\\
&&\quad\quad\quad\quad = \int d^4y'd^4z' H_1(x,y',z')S_F(y'-y)\Delta(z'-z), \\
&&i e^{i \partial_y \wedge\partial_{\hat y}} \langle 0|T(\psi(x)\bar\psi(y) C(\hat y)\bar C(z))|0\rangle|_{\hat y=y} \nonumber\\
&&\quad\quad\quad\quad = \int d^4x'd^4z' S_F(x-x')H_2(x',y,z')\Delta(z'-z),
\end{eqnarray}
where $S_F(x-x')$ and $\Delta(z'-z)$ are the fermion and ghost fields propagators, respectively, $\Gamma^\nu(x',y',z')$ is the vectorial proper vertex,
\begin{equation}
\Gamma^\nu(x',y',z') = \frac{\delta^3 \Gamma}{\delta\bar\psi_{cl}(x')\delta\psi_{cl}(y')\delta A_{cl\nu}(z')},
\end{equation}
\begin{equation}
H_1(x,y',z') = i \int d^4u\,d^4v\, e^{i \partial_x \wedge\partial_{\hat x}} S_F(x-u)\Delta(\hat x-v)\Gamma(u,y',v,z')\vert_{\hat x=x}
\end{equation}
and
\begin{equation}
H_2(x',y,z') = i \int d^4u\,d^4v\, \Gamma(x',u,v,z') e^{i \partial_y \wedge\partial_{\hat y}} S_F(u-y)\Delta(\hat y-v)\vert_{\hat y=y},
\end{equation}
in which
\begin{equation}
\Gamma(u,y',v,z') =\left. \frac{\delta^4 \Gamma}{\delta\bar\psi_{cl}(u)\delta\psi_{cl}(y')\delta\bar C_{cl}(v)\delta C_{cl}(z')}\right |
\end{equation}
is the fermion-ghost four-vertex.
In momentum space, Eq. (\ref{vertex}) reads
\begin{equation}\label{WTvertex}
\Gamma^\nu(k,p,q)q^\mu D_{\mu\nu}(-q) = -i \alpha e \left[S_{F}^{-1}(k)H_1(k,p,q)\Delta(q) - H_2(k,p,q)S_{F}^{-1}(p)\Delta(q) \right],
\end{equation}
where
\begin{equation}
H_1(k,p,q) = i \int \frac{d^4k'}{(2\pi)^4} e^{ik'\wedge k}S_F(k')\Delta(k-k')\Gamma(k',p,k-k',q)
\end{equation}
and
\begin{equation}
H_2(k,p,q) = i \int \frac{d^4p'}{(2\pi)^4} e^{-i p'\wedge p}\Gamma(k,p',p'-p,q)S_F(p')\Delta(p'-p),
\end{equation}
with $k=p+q$.
Similarly, we may determine $\Delta(q)$ from the Dyson-Schwinger equation,
\begin{equation}
\Delta(q) = \Delta_{(0)}(q) - \Delta_{(0)}(q) \Sigma_{C}(q) \Delta(q),
\end{equation}
where $\Delta_{(0)}(q)=\frac{i}{q^2}$ and $\Sigma_{C}(q)$ denotes the proper self-energy operator of the ghost field. Therefore, it is easy to verify that
\begin{equation}\label{ghost}
\Delta(q) = \frac{i}{q^2[1+b(q^2)]},
\end{equation}
in which the self-energy has been expressed as $i\Sigma_{C}(q)=q^2 b(q^2)$.
Thus, with the expressions (\ref{foton}) and (\ref{ghost}), we can rewrite the identity (\ref{WTvertex}) as follows
\begin{equation}\label{WTvm}
q_\mu\Gamma^\mu(k,p,q)[1+b(q^2)] = ie [S^{-1}_{F}(k)H_1(k,p,q) - H_2(k,p,q)S^{-1}_{F}(p)].
\end{equation}
By considering that energy-momentum conservation holds at the vertices $\Gamma^\mu(k,p,q)$ and $H(k,p,q)$, we can write
\begin{equation}
\Gamma^\mu(k,p,q) = ie(2\pi)^4 \delta^4(k-p-q)\tilde\Gamma^\mu(p,p+q)
\end{equation}
and
\begin{equation}
H(k,p,q) = (2\pi)^4 \delta^4(k-p-q)\tilde H(p,p+q).
\end{equation}
With this representation we may obtain from Eq. (\ref{WTvm}) that
\begin{equation}\label{2}
q_\mu\tilde\Gamma^\mu(p,p+q)[1+b(q^2)] = S_{F}^{-1}(p+q)\tilde H_1(p,p+q) - \tilde H_2(p,p+q)S_{F}^{-1}(p).
\end{equation}
\subsection{The triple photon vertex}
To obtain ST identity for the triple photon vertex, the proper part of $\langle 0|T(A_\mu A_\nu A_\lambda)|0\rangle$, we differentiate the functional equation (\ref{gfc}) with respect to $\zeta(x)$, $J^\nu(y)$ and $J^\lambda(z)$ and turn off all the sources. The result is
\begin{equation}
\frac1{\alpha e} \partial_x^\mu \frac{\delta^3 W}{\delta J^\mu(x)\delta J^\nu(y)\delta J^\lambda(z)}\left | = \frac{\delta^3 W}{\delta\zeta(x)\delta K^\nu(y)\delta J^\lambda(z)}\right | +\left. \frac{\delta^3 W}{\delta\zeta(x)\delta J^\nu(y)\delta K^\lambda(z)} \right | ,
\end{equation}
or in terms of the Green functions,
\begin{eqnarray}
-\frac1{\alpha} \partial_x^\mu \langle 0|T(A_\mu(x)A_\nu(y)A_\lambda(z))|0\rangle &=& \langle 0|T(\bar C(x)D^{AD}_\nu(y) C(y)A_\lambda(z))|0\rangle \nonumber\\
&& + \langle 0|T(\bar C(x)A_\nu(y)D^{AD}_\lambda(z)C(z))|0\rangle,
\end{eqnarray}
where $D^{AD}_\nu(y)$ denotes the covariant derivative in the adjoint representation, $D^{AD}_\nu(y) C(y)=\partial_{y\nu}C(y)-ie[A_\nu(y),C(y)]_\star$. Thus, we can rewrite the above expression as
\begin{eqnarray}\label{triplevertex}
&&-\frac1{\alpha} \partial_{x}^{\mu} \langle 0|T(A_\mu(x)A_\nu(y)A_\lambda(z))|0\rangle = \partial_{y\nu} \langle 0|T(\bar C(x) C(y)A_\lambda(z))|0\rangle \nonumber\\
&&+ 2e\sin(\partial_y\wedge\partial_{\hat y}) \langle 0|T(\bar C(x)A_\nu(y)C(\hat y)A_\lambda(z))|0\rangle +\partial_{z\lambda} \langle 0|T(\bar C(x)A_\nu(y)C(z))|0\rangle \nonumber\\
&&+ 2e\sin(\partial_z\wedge\partial_{\hat z}) \langle 0|T(\bar C(x)A_\nu(y)A_\lambda(z)C(\hat z))|0\rangle,
\end{eqnarray}
where, after the application of the differential operators, we must identify $\hat y$ and $\hat z$ respectively with $y$ and $z$.
These Green functions have the following one-particle irreducible decomposition:
\begin{eqnarray}
&&\langle 0|T(A_\mu(x)A_\nu(y)A_\lambda(z))|0\rangle \nonumber\\
&&\quad\quad\quad\quad = \int d^4x'd^4y'd^4z' D_{\mu\mu'}(x-x')D_{\nu\nu'}(y-y')D_{\lambda\lambda'}(z-z')\Gamma^{\mu'\nu'\lambda'}(x',y',z'), \\
&&\partial_{y\nu}\langle 0|T(\bar C(x) C(y)A_\lambda(z))|0\rangle \nonumber\\
&&\quad\quad\quad\quad = i\int d^4x'd^4y'd^4z' \Delta(x-x')\partial_{y\nu}\Delta(y-y')D_{\lambda\rho}(z-z')G^\rho(x',y',z'), \\
&& 2e\sin(\partial_y\wedge\partial_{\hat{y}}) \langle 0|T(\bar C(x)A_\nu(y)C(\hat{y})A_\lambda(z))|0\rangle \nonumber\\
&&\quad\quad\quad\quad = -i\int d^4x'd^4z' \Delta(x'-x) G_\nu^{\;\;\lambda'}(x',y,z')D_{\lambda'\lambda}(z'-z),\\
&&\partial_{z\lambda} \langle 0|T(\bar C(x)A_\nu(y)C(z))|0\rangle \nonumber\\
&&\quad\quad\quad\quad = i\int d^4x'd^4y'd^4z' \Delta(x'-x)D_{\nu\rho}(y-y')\partial_{z\lambda}\Delta(z-z')G^\rho(x',z',y'),\\
&&2e\sin(\partial_z\wedge\partial_{\hat{z}}) \langle 0|T(\bar C(x)A_\nu(y)A_\lambda(z)C(\hat{z}))|0\rangle \nonumber\\
&&\quad\quad\quad\quad = -i\int d^4x'd^4y' \Delta(x'-x) {G_\lambda}^{\nu'}(x',z,y')D_{\nu'\nu}(y'-y),
\end{eqnarray}
where $\Gamma^{\mu'\nu'\lambda'}(x',y',z')$ and $G^\rho(x',z',y')$ are the triple gauge and the ghost-gauge vertices respectively,
\begin{equation}\label{AAA}
\Gamma^{\mu'\nu'\lambda'}(x',y',z') = \frac{\delta^3 \Gamma}{\delta A_{cl\,\mu'}(x')\delta A_{cl\,\nu'}(y')\delta A_{cl\,\lambda'}(z')}
\end{equation}
and
\begin{equation}
G^\rho(x',z',y') = \frac{\delta^3 \Gamma}{\delta C_{cl}(x')\delta\bar C_{cl}(y')\delta A_{cl\,\rho}(z')}.
\end{equation}
Also
\begin{equation}\label{G}
{G_\nu}^{\lambda'}(x',y,z') = -2e\int d^4u\,d^4v\, \sin(\partial_y\wedge\partial_{\hat{y}}) \Delta(y-u)D_\nu^{\;\;\nu'}(\hat{y}-v)\Gamma_{\nu'}^{\;\;\;\lambda'}(x',u,v,z'),
\end{equation}
in which
\begin{equation}
\Gamma_{\nu'}^{\;\;\;\lambda'}(x',u,v,z') = \frac{\delta^4 \Gamma}{\delta C_{cl}(x')\delta\bar C_{cl}(u)\delta A^{\nu'}_{cl}(v)\delta A_{cl\,\lambda'}(z')}.
\end{equation}
In momentum space the Eq.~(\ref{triplevertex}) reads
\begin{eqnarray}\label{3}
p^\mu \Gamma_{\mu\nu\lambda}(p,q,k)[1+b(p^2)] &=& G_{\lambda\nu'}(p,q,k)\{(q^2g^{\nu'}_{\;\;\;\nu}-q^{\nu'}q_{\nu})[1+\Pi_\mathrm{T}(q^2)] + \Pi_\theta(q^2)\tilde{q}^{\nu'}\tilde{q}_\nu\} \nonumber\\
&+& \!\!\! G_{\nu\lambda'}(p,k,q)\{(k^2g^{\lambda'}_{\;\;\;\lambda}-k^{\lambda'}k_{\lambda})[1+\Pi_\mathrm{T}(k^2)] + \Pi_\theta(k^2)\tilde{k}^{\lambda'}\tilde{k}_\lambda\}
\end{eqnarray}
with
\begin{equation}
G_{\lambda\nu'}(p,q,k)=-2e\int \frac{d^4q'}{(2\pi)^4} \sin(q'\wedge q)i\Delta(q')iD_{\nu'}^{\;\;\;\alpha}(q-q')\Gamma_{\alpha\lambda}(p,q',q-q',k),
\end{equation}
where we have used the ghost propagator (\ref{ghost}) and the inverse of the photon propagator
\begin{equation}
iD^{-1}_{\mu\nu}(q) = -i \left\{\left(q^2g_{\mu\nu}-q_\mu q_\nu\right)[1+\Pi_\mathrm{T}(q^2)] - \frac{q_\mu q_\nu}{\alpha} - \Pi_\theta(q^2) \tilde{q}_\mu\tilde{q}_\nu\right\},
\end{equation}
which satisfy
\begin{equation}
q^\mu D^{-1}_{\mu\nu}(q) = -\frac{1}{\alpha} q^2 q_\nu.
\end{equation}
Using the energy-momentum conservation at the vertices $\Gamma_{\mu\nu\lambda}(p,q,k)$ and $G_{\nu'\lambda}(p,q,k)$, so that
\begin{equation}
\Gamma_{\mu\nu\lambda}(p,q,k) = (2\pi)^4 \delta^4(p+q+k)\tilde\Gamma_{\mu\nu\lambda}(p,q,-p-q)
\end{equation}
and
\begin{equation}
G_{\lambda\nu'}(p,q,k) = (2\pi)^4 \delta^4(p+q+k) \tilde G_{\lambda\nu'}(p,q,-p-q),
\end{equation}
we get
\begin{eqnarray}\label{3}
p^\mu \tilde\Gamma_{\mu\nu\lambda}(p,q,k)[1+b(p^2)] &=&
\tilde G_{\lambda\nu'}(p,q,k)\{(q^2{g^{\nu'}}_{\nu}-q^{\nu'}q_{\nu})[1+\Pi_\mathrm{T}(q^2)] + \Pi_\theta(q^2)\tilde{q}^{\nu'}\tilde{q}_\nu\} \\
&&+\tilde G_{\nu\lambda'}(p,k,q)\{[k^2{g^{\lambda'}}_{\lambda}-k^{\lambda'}k_{\lambda}][1+\Pi_\mathrm{T}(k^2)] + \Pi_\theta(k^2)\tilde{k}^{\lambda'}\tilde{k}_\lambda\}. \nonumber
\end{eqnarray}
\section{Slavnov-Taylor identities at one-loop: explicit calculations}
\subsection{The vectorial vertex function}
The ST identities derived previously are valid only in a formal way since the radiative corrections contain ultraviolet divergences.
To eliminate these divergences counterterms must be introduced so that
the action for noncommutative QED$_4$ becomes
\begin{eqnarray}
S&=& \int d^4 x \left [-\frac{Z_3}{4} G_{\mu\nu}G^{\mu\nu}- \frac{ieZ_1}{2} [A_\mu,A_\nu]_\star G^{\mu\nu}
+ \frac{e^2 Z_4}{4} [A_\mu,A_\nu]_\star [A^\mu,A^\nu]_\star+ Z_{2}\bar\psi i \slash \!\!\!{\partial}\psi\right. \nonumber\\
&&\!\!- (m+\delta m) \bar \psi \psi+ e\left. Z_{1F} \bar \psi\star \slash\!\!\!\! A \psi - \frac{1}{2\alpha} (\partial_\mu A^\mu)^2 + \tilde Z_3\partial_\mu \bar C \partial^\mu C - ie\tilde Z_1 \partial_\mu \bar C \star [A^\mu,C]_\star \right ],
\end{eqnarray}
where $G_{\mu\nu}=\partial_\mu A_\nu - \partial_\nu A_\mu$.
We begin by considering the one-loop contributions to the vectorial vertex function. Writing the $\tilde\Gamma^\mu(p,p+q)$, $S^{-1}(p+q)$ and $\tilde H(p,p+q)$ expansions as
\begin{equation}
\tilde\Gamma^\mu(p,p+q) = Z_{1F}\gamma^\mu e^{ip\wedge q} + \Lambda^\mu(p,p+q)e^{ip\wedge q},
\end{equation}
\begin{equation}
\tilde H_i(p,p+q) = \left[Z_{i+4} e^{ip\wedge q} + B_i(p,p+q)e^{ip\wedge q}\right] \qquad \mbox{for $i=1,2$}
\end{equation}
and
\begin{equation}
S_{F}^{-1}(p) = \left[Z_2\slashed{p}-m-\delta m - \Sigma(p)\right],
\end{equation}
where $\Lambda^\mu(p,p+q)$, $\Sigma(p)$ and $B(p,p+q)$ are the one-loop contributions. Thus, the ST identity for the vectorial vertex (\ref{2}), in the tree approximation, becomes
\begin{eqnarray}\label{tree}
Z_{1F}\tilde Z_3\slashed{q}e^{ip\wedge q} = [Z_2 Z_5 (\slashed{p} + \slashed{q}) - (m+\delta m) Z_5] e^{ip\wedge q} - [Z_2 Z_6 \slashed{p} - (m+\delta m) Z_6] e^{ip\wedge q},
\end{eqnarray}
so that the validity of the ST identity requires that
\begin{equation}
Z_5=Z_6\qquad \mbox{and}\qquad \tilde Z_3/Z_5= Z_2/Z_{1F} .
\end{equation}
For the one-loop approximation, we have
\begin{eqnarray}\label{loop}
q_\mu \Lambda^\mu_a(p,p+q) + q_\mu \Lambda^\mu_b(p,p+q) + \slashed q b(q^2) &=& \Sigma(p) - \Sigma(p+q) + (\slashed{p}+\slashed{q}-m)B_1(p,p+q) \nonumber\\
&& - B_2(p,p+q)(\slashed{p}-m).
\end{eqnarray}
The diagrams representing these contributions are given by: (from now on, we restrict ourselves to the Feynman gauge, $\alpha=1$)
\begin{eqnarray}
1.\raisebox{-0.4cm}{\incps{vertexlambdaa.eps}{-1.5cm}{-0.5cm}{1.5cm}{1.5cm}}
&=& -ie \Lambda^\mu_a(p,p+q) e^{i p\wedge q} \\ \nonumber\\ \nonumber\\
&=& \int \frac{d^4l}{(2\pi)^4} \frac{-ig_{\alpha\beta}}{l^2} (-ie\gamma^\alpha) iS_0(p+q-l) (-ie\gamma^\mu) iS_0(p-l) (-ie\gamma^\beta) e^{-2i l\wedge q}e^{i p\wedge q}, \nonumber
\end{eqnarray}
with
\begin{equation}
iS_0(p)=\frac{i}{\slashed{p}-m}.
\end{equation}
This contribution is entirely nonplanar and using $ \slashed{q}= (\slashed{q}+\slashed p-\slashed l-m)-(\slashed p-\slashed l-m)$ can be shown to satisfy
\begin{equation}
q_\mu \Lambda^\mu_a(p,p+q) = \Sigma_\mathrm{np}(p) - \Sigma_\mathrm{np}(p+q),
\end{equation}
where the nonplanar fermion self-energy is
\begin{equation}\label{sigma}
-i\Sigma_\mathrm{np}(p) = \int \frac{d^4l}{(2\pi)^4} \frac{-ig_{\alpha\beta}}{l^2}(-ie\gamma^\alpha)iS_0(p-l)(-ie\gamma^\beta)e^{-2il\wedge q}.
\end{equation}
\begin{eqnarray}\label{L2}
2.\raisebox{-0.4cm}{\incps{vertexlambdab.eps}{-1.5cm}{-0.5cm}{1.5cm}{1.5cm}}
&=& -ie \Lambda^\mu_b(p,p+q) e^{i p\wedge q} \\ \nonumber\\ \nonumber\\
&=& \int \frac{d^4l}{(2\pi)^4} \frac{-ig_{\alpha\beta}}{(p+q-l)^2}\frac{-ig_{\lambda\rho}}{(p-l)^2} (-ie\gamma^\alpha) iS_0(l) (-ie\gamma^\lambda) \nonumber\\
&& \times (2e)\gamma^{\mu\beta\rho}(q,-p-q+l,p-l)\frac1{2i}\left(1-e^{2i l\wedge q}e^{-2i p\wedge q}\right)e^{i p\wedge q}, \nonumber
\end{eqnarray}
whose planar part logarithimically diverges. In fact its pole part (PP) is given by
\begin{equation}
\mathrm{PP}[-ie \Lambda^\mu_b(p,p+q)] = \frac{3}{16\pi^{2}}\frac{1}{\epsilon}\gamma^\mu
\end{equation}
\noindent
so that, in the minimal dimensional regularization scheme,
\begin{equation}
Z_{1F} = 1- \frac{3}{16\pi^{2}}\frac{1}{\epsilon},
\end{equation}
which agrees with the result of previous calculation \cite{Hayakawa}. Contracting $q_\mu$ in the expression~(\ref{L2}), we get
\begin{eqnarray}
q_\mu \Lambda^\mu_b(p,p+q) &=& \int \frac{d^4l}{(2\pi)^4} \frac{-ig_{\alpha\beta}}{(p+q-l)^2}\frac{-ig_{\lambda\rho}}{(p-l)^2} (-ie\gamma^\alpha) iS_0(l) (-ie\gamma^\lambda) \left(1-e^{2i l\wedge q}e^{-2i p\wedge q}\right) \nonumber\\
&& \times \left\{(p+q-l)^\rho q^\beta + (p-l)^\beta q^\rho - [(p+q-l)\cdot q + (p-l)\cdot q] g^{\beta\rho}\right\},
\end{eqnarray}
which may be further simplified using
\begin{eqnarray}
(p+q-l)^\rho q^\beta + (p-l)^\beta q^\rho &=& (p+q-l)^\rho (p+q-l)^\beta - (p-l)^\rho (p-l)^\beta,\;\mathrm{and} \nonumber\\
(p+q-l)\cdot q + (p-l)\cdot q &=& (p+q-l)^2 - (p-l)^2,
\end{eqnarray}
to yield
\begin{eqnarray}
q_\mu \Lambda^\mu_b(p,p+q) &=& \Sigma(p) - \Sigma(p+q) - \Sigma_\mathrm{np}(p) + \Sigma_\mathrm{np}(p+q) \\
&& + \int \frac{d^4l}{(2\pi)^4} \frac{i}{(p+q-l)^2}\frac{i}{(p-l)^2}(ie)(\slashed{p}+\slashed{q}-\slashed{l})iS_0(l)(ie)(\slashed{p}+\slashed{q}-\slashed{l})\left(1-e^{2i l\wedge q}e^{-2i p\wedge q}\right) \nonumber\\
&& - \int \frac{d^4l}{(2\pi)^4} \frac{i}{(p+q-l)^2}\frac{i}{(p-l)^2}(ie)(\slashed{p}-\slashed{l})iS_0(l)(ie)(\slashed{p}-\slashed{l})\left(1-e^{2i l\wedge q}e^{-2i p\wedge q}\right), \nonumber
\end{eqnarray}
where $\Sigma$ is similar to $\Sigma_\mathrm{np}$ of Eq.~(\ref{sigma}), but without the phase factor.
For the last term in left-hand side of Eq.~(\ref{loop}), we get
\begin{eqnarray}\label{gse}
\raisebox{-0.4cm}{\incps{ghostself.eps}{-1.5cm}{-0.5cm}{2.0cm}{1.5cm}}
&=& i q^2 b(q^2) \\ \nonumber\\ \nonumber\\
&=& - \int \frac{d^4l}{(2\pi)^4} \frac{-ig_{\alpha\beta}}{l^2} \frac{i}{(q-l)^2} (2e)(q-l)^\alpha(2e)q^\beta \frac12 \left(1- e^{2il\wedge q}\right). \nonumber
\end{eqnarray}
Then,
\begin{eqnarray}
\slashed{q}b(q^2) &=& \frac{2i}{q^2} \int \frac{d^4l}{(2\pi)^4} \frac{i}{l^2}\frac{i}{(q-l)^2}(ie)\slashed{q}(ie)(q-l)\cdot q \left(1 - e^{2il\wedge q}\right) \\
&\overbrace{=}^{l \rightarrow l-p}& \frac{2i}{q^2} \int \frac{d^4l}{(2\pi)^4} \frac{i}{(p-l)^2}\frac{i}{(p+q-l)^2}(ie)\slashed{q}(ie)(p+q-l)\cdot q \left(1 - e^{2il\wedge q}e^{-2ip\wedge q}\right). \nonumber
\end{eqnarray}
Since
\begin{equation}
2(p+q-l)\cdot q = (p+q-l)^2 - (p-l)^2 + q^2,
\end{equation}
we find
\begin{equation}
\slashed{q}b(q^2) = i \int \frac{d^4l}{(2\pi)^4} \frac{i}{(p-l)^2}\frac{i}{(p+q-l)^2}(ie)\slashed{q}(ie)\left(1 - e^{2il\wedge q}e^{-2ip\wedge q}\right).
\end{equation}
Hence, for the left-hand side of Eq. (\ref{loop}), we find
\begin{eqnarray}\label{rhs}
&& q_\mu \Lambda^\mu_a(p,p+q) + q_\mu \Lambda^\mu_b(p,p+q)\; + \slashed{q}b(q^2) = \Sigma(p) - \Sigma(p+q) \nonumber\\
&& + \int \frac{d^4l}{(2\pi)^4} \frac{i}{(p+q-l)^2}\frac{i}{(p-l)^2}(ie)(\slashed{p}+\slashed{q}-\slashed{l})iS_0(l)(ie)(\slashed{p}+\slashed{q}-\slashed{l})\left(1-e^{2i l\wedge q}e^{-2i p\wedge q}\right) \nonumber\\
&& - \int \frac{d^4l}{(2\pi)^4} \frac{i}{(p+q-l)^2}\frac{i}{(p-l)^2}(ie)(\slashed{p}-\slashed{l})iS_0(l)(ie)(\slashed{p}-\slashed{l})\left(1-e^{2i l\wedge q}e^{-2i p\wedge q}\right) \nonumber\\
&& + i \int \frac{d^4l}{(2\pi)^4} \frac{i}{(p-l)^2}\frac{i}{(p+q-l)^2}(ie)\slashed{q}(ie)\left(1 - e^{2il\wedge q}e^{-2ip\wedge q}\right).
\end{eqnarray}
Let us now consider the one-loop contributions to the right-hand side of Eq. (\ref{loop}). We have
\begin{eqnarray}
\raisebox{-0.4cm}{\incps{B1.eps}{-1.5cm}{-0.5cm}{1.2cm}{1.5cm}}
&=& B_1(p,p+q)e^{ip\wedge q} \\ \nonumber\\ \nonumber\\
&=& -\int \frac{d^4l}{(2\pi)^4} iS_0(l)\frac{i}{(p+q-l)^2}(-ie\gamma^\alpha)\frac{-ig_{\alpha\beta}}{(p-l)^2}(2e)(p+q-l)^\beta \nonumber\\
&& \times \frac{1}{2i}\left(1 - e^{2il\times q}e^{-2ip\times q}\right)e^{ip\times q}, \nonumber
\end{eqnarray}
yielding
\begin{equation}
Z_5= 1-\frac{e^2}{16\pi^2} \frac{1}{\epsilon}.
\end{equation}
Notice that in the Landau gauge $Z_5=1$ as it happens in ordinary commutative QCD \cite{marciano}. Now,
\begin{eqnarray}
(\slashed{p}\;+\slashed{q}-m)B_1(p,p+q) &=& \int \frac{d^4l}{(2\pi)^4} \frac{i}{(p+q-l)^2}\frac{i}{(p-l)^2}(ie)(\slashed{p}+\slashed{q}-m)iS_0(l)(ie)(\slashed{p}+\slashed{q}-\slashed{l})\nonumber\\
&&\times\left(1-e^{2i l\times q}e^{-2i p\times q}\right)
\end{eqnarray}
and making the substitution $(\slashed{p}+\slashed{q}-m) \rightarrow (\slashed{p}+\slashed{q}-\slashed{l}) + (\slashed{l} - m)$, we obtain
\begin{eqnarray}
(\slashed{p}\;+\slashed{q}-m)B_1(p,p+q) &=& \int \frac{d^4l}{(2\pi)^4} \frac{i}{(p+q-l)^2}\frac{i}{(p-l)^2}(ie)(\slashed{p}+\slashed{q}-\slashed{l})iS_0(l)(ie)(\slashed{p}+\slashed{q}-\slashed{l})\nonumber\\
&&\times\left(1-e^{2i l\wedge q}e^{-2i p\wedge q}\right) \nonumber\\
&& + i \int \frac{d^4l}{(2\pi)^4} \frac{i}{(p+q-l)^2}\frac{i}{(p-l)^2}(ie)(\slashed{p}+\slashed{q}-\slashed{l})(ie)\nonumber\\
&&\times\left(1-e^{2il\wedge q}e^{-2ip\wedge q}\right).
\end{eqnarray}
Similarly,
\begin{eqnarray}
\raisebox{-0.4cm}{\incps{B2.eps}{-1.5cm}{-0.5cm}{1.7cm}{1.5cm}}
&=& B_2(p,p+q)e^{ip\wedge q} \\ \nonumber\\ \nonumber\\
&=& \int \frac{d^4l}{(2\pi)^4} (-ie\gamma^\alpha) \frac{-ig_{\alpha\beta}}{(p+q-l)^2}(2e)(l-p)^\beta \frac{i}{(p-l)^2} iS_0(l) \nonumber\\
&&\times\frac{1}{2i}\left(1 - e^{2il\times q}e^{-2ip\times q}\right)e^{ip\times q}. \nonumber
\end{eqnarray}
Therefore,
\begin{eqnarray}
B_2(p,p+q)(\slashed{p}-m) &=& \int \frac{d^4l}{(2\pi)^4} \frac{i}{(p+q-l)^2}\frac{i}{(p-l)^2}(ie)(\slashed{p}-\slashed{l})iS_0(l)(ie)(\slashed{p}-\slashed{l}) \nonumber\\
&&\times\left(1-e^{2il\wedge q}e^{-2ip\wedge q}\right) \nonumber\\
&& + i \int \frac{d^4l}{(2\pi)^4} \frac{i}{(p+q-l)^2}\frac{i}{(p-l)^2} (ie)(\slashed{p}-\slashed{l})(ie) \nonumber\\
&&\times\left(1-e^{2il\times q}e^{-2ip\times q}\right),
\end{eqnarray}
where we also have done the replacement $(\slashed{p}-m) \rightarrow (\slashed{p}-\slashed{l}) + (\slashed{l} - m)$. Finally, summing the above results,
\begin{eqnarray}
&&(\slashed{p}+\slashed{q}-m)B_1(p,p+q) - B_2(p,p+q)(\slashed{p}-m) = \nonumber\\
&& + \int \frac{d^4l}{(2\pi)^4} \frac{i}{(p+q-l)^2}\frac{i}{(p-l)^2}(ie)(\slashed{p}+\slashed{q}-\slashed{l})iS_0(l)(ie)(\slashed{p}+\slashed{q}-\slashed{l})\left(1-e^{2i l\wedge q}e^{-2i p\wedge q}\right) \nonumber\\
&& - \int \frac{d^4l}{(2\pi)^4} \frac{i}{(p+q-l)^2}\frac{i}{(p-l)^2}(ie)(\slashed{p}-\slashed{l})iS_0(l)(ie)(\slashed{p}-\slashed{l})\left(1-e^{2i l\wedge q}e^{-2i p\wedge q}\right) \nonumber\\
&& + i \int \frac{d^4l}{(2\pi)^4} \frac{i}{(p-l)^2}\frac{i}{(p+q-l)^2}(ie)\slashed{q}(ie)\left(1 - e^{2il\wedge q}e^{-2ip\wedge q}\right).
\end{eqnarray}
Thus, the left-hand side of Eq.~(\ref{loop}) is identical to the right-hand side as we can see from Eq.~(\ref{rhs}). Therefore, the ST identity for the vectorial vertex is satisfied at one-loop.
\subsection{The triple photon vertex}
Writing the expansions for $\tilde\Gamma^{\mu\nu\lambda}(p,q,k)$ and $\tilde G^\lambda_{\;\;\nu'}(p,q,k)$, defined in~(\ref{AAA}) and~(\ref{G}), as
\begin{equation}
\tilde\Gamma^{\mu\nu\lambda}(p,q,k) = 2eZ_1\sin(p\wedge q)\gamma^{\mu\nu\lambda}(p,q,k) + 2e\sin(p\wedge q)\Lambda^{\mu\nu\lambda}(p,q,k),
\end{equation}
\begin{equation}
\tilde G^\lambda_{\;\;\nu'}(p,q,k) = -2e\tilde Z_1\sin(p\wedge q){g^\lambda}_{\nu'} + 2e\sin(p\wedge q){B^\lambda}_{\nu'}(p,q,k),
\end{equation}
we obtain the ST identity (\ref{3}) for the triple photon vertex, in the tree approximation,
\begin{equation}
Z_1\tilde Z_3\left[(k^2 g^{\lambda\nu} - k^\lambda k^\nu)-(q^2 g^{\nu\lambda} - q^\nu q^\lambda)\right]=\tilde Z_1Z_3\left[(k^2 g^{\lambda\nu} - k^\lambda k^\nu)-(q^2 g^{\nu\lambda} - q^\nu q^\lambda)\right],
\end{equation}
which requires that
\begin{equation}
\tilde Z_3/\tilde Z_1=Z_3/Z_1.
\end{equation}
On the other hand, the one-loop approximation is given by
\begin{eqnarray}\label{ITF}
&& p_\mu \gamma^{\mu\nu\lambda}(p,q,k)b(p^2) + p_\mu \Lambda^{\mu\nu\lambda}(p,q,k) = \Pi^{\nu\lambda}(q) - \Pi^{\lambda\nu}(k) \nonumber\\
&&+ B^\lambda_{\;\;\nu'}(p,q,k)(q^2 g^{\nu\nu'} - q^\nu q^{\nu'}) + B^\nu_{\;\;\lambda'}(p,k,q)(k^2 g^{\lambda\lambda'} - k^\lambda k^{\lambda'}),
\end{eqnarray}
where we have introduced the photon self-energy $\Pi^{\nu\nu'}(q)\equiv\Pi_\mathrm{T}(q^2)(q^2g^{\nu\nu'}-q^{\nu}q^{\nu'}) + \Pi_\theta(q^2)\tilde{q}^{\nu}\tilde{q}^{\nu'}$.
The contributions with a fermion loop in the left and right-hand sides of Eq.~(\ref{ITF}) are directly identified when we consider the diagrams:
\begin{eqnarray}
\raisebox{-0.4cm}{\incps{triplefermion.eps}{-1.8cm}{-0.5cm}{2cm}{1.8cm}} &=& 2e\sin(p\wedge q)\Lambda_{a1}^{\mu\nu\lambda}(p,q,k) \\ \nonumber\\ \nonumber\\
&=& -\int \frac{d^4l}{(2\pi)^4} \mathrm{tr}(-ie\gamma^\nu)iS_0(q+l)(-ie\gamma^\mu)iS_0(p+q+l)\nonumber\\
&&\times(-ie\gamma^\lambda)iS_0(l)e^{-i p\wedge q}, \nonumber \\
\raisebox{-0.4cm}{\incps{triplefermioni.eps}{-1.8cm}{-0.5cm}{2cm}{1.8cm}} &=& 2e\sin(p\wedge q)\Lambda_{a2}^{\mu\nu\lambda}(p,q,k) \\ \nonumber\\ \nonumber\\
&=& -\int \frac{d^4l}{(2\pi)^4} \mathrm{tr}\,iS_0(-l)(-ie\gamma^\lambda)iS_0(-p-q-l)\nonumber\\
&&\times(-ie\gamma^\mu)iS_0(-q-l)(-ie\gamma^\nu)e^{i p\wedge q}. \nonumber
\end{eqnarray}
These diagrams are different only in the circulation of the momentum integration. Since $C\gamma^\mu C^{-1}=-\gamma^{\mathrm{T}\mu}$ and $CS_0(l)C^{-1}=S_0^\mathrm{T}(-l)$, we can rewrite the above expression as
\begin{eqnarray}
2e\sin(p\wedge q)\Lambda_{a2}^{\mu\nu\lambda}(p,q,k)
&=& e^3 \int \frac{d^4l}{(2\pi)^4} \mathrm{tr}\,S_0^\mathrm{T}(l)\gamma^{\mathrm{T}\lambda}S_0^\mathrm{T}(p+q+l)\gamma^{\mathrm{T}\mu}S_0(q+l)\gamma^{\mathrm{T}\nu} e^{i p\wedge q} \nonumber\\
&=& e^3 \int \frac{d^4l}{(2\pi)^4} \mathrm{tr}[\gamma^\nu S_0(q+l)\gamma^\mu S_0(p+q+l)\gamma^\lambda S_0(l)]^\mathrm{T} e^{i p\wedge q} \nonumber\\
&=& e^3 \int \frac{d^4l}{(2\pi)^4} \mathrm{tr}\,\gamma^\nu S_0(q+l)\gamma^\mu S_0(p+q+l)\gamma^\lambda S_0(l) e^{i p\wedge q}.
\end{eqnarray}
Thus, by summing the two diagrams, we obtain
\begin{eqnarray}
2e\sin(p\wedge q)\Lambda_{a}^{\mu\nu\lambda}(p,q,k) &=& 2e \sin(p\wedge q)\left[\Lambda_{a1}^{\mu\nu\lambda}(p,q,k)+\Lambda_{a2}^{\mu\nu\lambda}(p,q,k)\right] \\
&=& 2ie^3 \int \frac{d^4l}{(2\pi)^4} \mathrm{tr}\,\gamma^\nu S_0(q+l)\gamma^\mu S_0(p+q+l)\gamma^\lambda S_0(l) \sin(p\wedge q), \nonumber
\end{eqnarray}
i.e.,
\begin{equation}
\Lambda_{a}^{\mu\nu\lambda}(p,q,k) = ie^2 \int \frac{d^4l}{(2\pi)^4} \mathrm{tr}\,\gamma^\nu S_0(q+l)\gamma^\mu S_0(p+q+l)\gamma^\lambda S_0(l).
\end{equation}
Therefore, using also $(\slashed{p}+\slashed{q}+\slashed{l}-m)-(\slashed{q}+\slashed{l}-m)$, we get
\begin{equation}
p_\mu\Lambda_{a}^{\mu\nu\lambda}(p,q,k) = \Pi^{\nu\lambda}_a(q) - \Pi^{\lambda\nu}_a(k),
\end{equation}
where
\begin{equation}
i\Pi^{\nu\lambda}_a(q) = -\int \frac{d^4l}{(2\pi)^4} \mathrm{tr}(-ie\gamma^\nu)iS_0(q+l)(-ie\gamma^\lambda)iS_0(l)
\end{equation}
is the photon self-energy, with a fermion loop.
From now on, differently for the previous calculations, the contributions to the ST identity turns out to be very involved and a complete verification is unfeasible. In this situation we restrict ourselves in to verify the matching of the divergent parts of the two sides of Eq.~(\ref{ITF}).
The diagram for the ghost self-energy has already been considered in (\ref{gse}) and its PP is given by
\begin{equation}
\mathrm{PP}[p_\mu \gamma^{\mu\nu\lambda}(p,q,k)b(p^2)] = \frac{e^2}{16\pi^2}\frac{1}{\epsilon}[(k^2 g^{\lambda\nu} - k^\lambda k^\nu) - (q^2 g^{\nu\lambda} - q^\nu q^\lambda)].
\end{equation}
In the sequel we consider the diagrams $\Lambda^{\mu\nu\lambda}$ of the left-hand side of~(\ref{ITF}), with ghost loop,
\begin{eqnarray}
\raisebox{-0.4cm}{\incps{tripleghost.eps}{-1.8cm}{-0.5cm}{2cm}{1.8cm}} &=& 2e\sin(p\wedge q)\Lambda_{b1}^{\mu\nu\lambda}(p,q,k) \\ \nonumber\\ \nonumber\\
&=& -i^3(2e)^3 \int \frac{d^4l}{(2\pi)^4} \frac{l^\lambda(p+l)^\mu(p+q+l)^\nu}{l^2(p+l)^2(p+q+l)^2} \nonumber\\\
&&\times \sin(l\wedge p)\sin(l\wedge p + l\wedge q)\sin(l\wedge q + p\wedge q), \nonumber\\
\raisebox{-0.4cm}{\incps{tripleghosti.eps}{-1.8cm}{-0.5cm}{2cm}{1.8cm}} &=& 2e\sin(p\wedge q)\Lambda_{b2}^{\mu\nu\lambda}(p,q,k) \\ \nonumber\\ \nonumber\\
&=& -i^3(2e)^3 \int \frac{d^4l}{(2\pi)^4} \frac{(-p-q-l)^\lambda(-p-l)^\nu(-l)^\mu}{l^2(p+l)^2(p+q+l)^2} \nonumber\\
&&\times \sin(l\wedge p)\sin(-l\wedge p - l\wedge q)\sin(l\wedge q + p\wedge q), \nonumber
\end{eqnarray}
and photon loop,
\begin{eqnarray}
\raisebox{-0.4cm}{\incps{triplephoton.eps}{-1.8cm}{-0.5cm}{2cm}{1.8cm}} &=& 2e\sin(p\wedge q)\Lambda_{c}^{\mu\nu\lambda}(p,q,k) \\ \nonumber\\ \nonumber\\
&=& i^3(2e)^3 \int \frac{d^4l}{(2\pi)^4} \frac{\gamma^{\alpha\lambda\beta}(p+q+l,-p-q,-l)\gamma^{\beta\mu\rho}(l,p,-p-l)}{l^2(p+l)^2(p+q+l)^2} \nonumber\\
&& \times \gamma^{\rho\nu\alpha}(l+p,q,-p-q-l)\sin(l\wedge p)\sin(-l\wedge p - l\wedge q)\sin(l\wedge q + p\wedge q), \nonumber
\end{eqnarray}
that have the same phase factors and, therefore, can be calculated analogously. Their PP contributions are
\begin{eqnarray}
{\rm PP}[p_\mu\Lambda_{b1,b2,c}^{\mu\nu\lambda}(p,q,k)]&=& -\frac{19e^2}{96\pi^2}\frac{1}{\epsilon}[(k^2 g^{\lambda\nu} - k^\lambda k^\nu) - (q^2 g^{\nu\lambda} - q^\nu q^\lambda)].
\end{eqnarray}
The remain diagrams for $\Lambda^{\mu\nu\lambda}$, are given by
\begin{eqnarray}
\raisebox{-0.4cm}{\incps{triplephotonf1.eps}{-1.8cm}{-0.5cm}{1.6cm}{2cm}} &=& 2e\sin(p\wedge q)\Lambda_{d1}^{\mu\nu\lambda}(p,q,k) \\ \nonumber\\ \nonumber\\
&=& \frac12 (-i)^2(2e)(-4ie^2) \int \frac{d^4l}{(2\pi)^4} \frac{(l-p)^\lambda g^{\mu\nu}+({g^\alpha}_\alpha-2)(2l+p)^\mu g^{\nu\lambda}+(l+2p)^\nu g^{\mu\lambda}}{l^2(p+l)^2} \nonumber\\
&& \times \sin(l\wedge p)\sin(-l\wedge p - l\wedge q)\sin(l\wedge q + p\wedge q) \nonumber\\
&& + \frac12 (-i)^2(2e)(-4ie^2) \int \frac{d^4l}{(2\pi)^4} \frac{(l+2p)^\lambda g^{\mu\nu}+({g^\alpha}_\alpha-2)(2l+p)^\mu g^{\nu\lambda}+(l-p)^\nu g^{\mu\lambda}}{l^2(p+l)^2} \nonumber\\
&& \times \sin(l\wedge p)\sin(l\wedge q)\sin(-l\wedge p -l\wedge q - p\wedge q) \nonumber\\
&& + \frac12 (-i)^2(2e)(-4ie^2) \int \frac{d^4l}{(2\pi)^4} \frac{3p^\lambda g^{\mu\nu}-3p^\nu g^{\mu\lambda}}{l^2(p+l)^2} \sin^2(l\wedge p) \sin(p\wedge q), \nonumber\\
\raisebox{-0.4cm}{\incps{triplephotonf2.eps}{-1.8cm}{-0.5cm}{1.6cm}{2cm}} &=& 2e\sin(p\wedge q)\Lambda_{d2}^{\mu\nu\lambda}(p,q,k) \\ \nonumber\\ \nonumber\\
&=& \frac12 (-i)^2(2e)(-4ie^2) \int \frac{d^4l}{(2\pi)^4} \frac{(l-q)^\lambda g^{\mu\nu}+(l+2q)^\mu g^{\nu\lambda}+({g^\alpha}_\alpha-2)(2l+q)^\nu g^{\mu\lambda}}{l^2(q+l)^2} \nonumber\\
&& \times \sin(l\wedge q)\sin(-l\wedge p - l\wedge q)\sin(l\wedge p - p\wedge q) \nonumber\\
&& + \frac12 (-i)^2(2e)(-4ie^2) \int \frac{d^4l}{(2\pi)^4} \frac{(l+2q)^\lambda g^{\mu\nu}+(l-q)^\mu g^{\nu\lambda}+({g^\alpha}_\alpha-2)(2l+q)^\nu g^{\mu\lambda}}{l^2(q+l)^2} \nonumber\\
&& \times \sin(l\wedge p)\sin(l\wedge q)\sin(-l\wedge p -l\wedge q + p\wedge q) \nonumber\\
&& + \frac12 (-i)^2(2e)(-4ie^2) \int \frac{d^4l}{(2\pi)^4} \frac{3q^\mu g^{\nu\lambda}-3q^\lambda g^{\mu\nu}}{l^2(q+l)^2} \sin^2(l\wedge q) \sin(p\wedge q),\quad\mathrm{and} \nonumber\\
\raisebox{-0.4cm}{\incps{triplephotonf3.eps}{-1.8cm}{-0.5cm}{1.6cm}{2cm}} &=& 2e\sin(p\wedge q)\Lambda_{d3}^{\mu\nu\lambda}(p,q,k) \\ \nonumber\\ \nonumber\\
&=& \frac12 (-i)^2(2e)(-4ie^2) \int \frac{d^4l}{(2\pi)^4} \frac{-({g^\alpha}_\alpha-2)(2l-k)^\lambda g^{\mu\nu}-(l+k)^\mu g^{\nu\lambda}-(l-2k)^\nu g^{\mu\lambda}}{l^2(p+q+l)^2} \nonumber\\
&& \times \sin(l\wedge p)\sin(-l\wedge p - l\wedge q)\sin(-l\wedge q - p\wedge q) \nonumber\\
&& + \frac12 (-i)^2(2e)(-4ie^2) \int \frac{d^4l}{(2\pi)^4} \frac{-({g^\alpha}_\alpha-2)(2l-k)^\lambda g^{\mu\nu}-(l-2k)^\mu g^{\nu\lambda}-(l+k)^\nu g^{\mu\lambda}}{l^2(p+q+l)^2} \nonumber\\
&& \times \sin(l\wedge q)\sin(-l\wedge p-l\wedge q)\sin(-l\wedge p + p\wedge q) \nonumber\\
&& + \frac12 (-i)^2(2e)(-4ie^2) \int \frac{d^4l}{(2\pi)^4} \frac{3k^\nu g^{\mu\lambda}-3k^\mu g^{\lambda\nu}}{l^2(p+q+l)^2} \sin^2(l\wedge p + l\wedge q) \sin(p\wedge q), \nonumber
\end{eqnarray}
where their PP contributions take the form
\begin{eqnarray}
{\rm PP}[p_\mu\Lambda_{d1,d2,d3}^{\mu\nu\lambda}(p,q,k)]&=& \frac{9e^2}{32\pi^2}\frac{1}{\epsilon}[(k^2 g^{\lambda\nu} - k^\lambda k^\nu) - (q^2 g^{\nu\lambda} - q^\nu q^\lambda)].
\end{eqnarray}
Therefore, the sum of all PP contributions of the left-hand side of Eq.~(\ref{ITF}) becomes
\begin{eqnarray}\label{sumL}
&& {\rm PP}[p_\mu(\gamma^{\mu\nu\lambda}(p,q,k)b(p^2)+\Lambda_{b1,b2,c,d1,d2,d3}^{\mu\nu\lambda}(p,q,k))] \nonumber\\
&&\quad\quad\quad\quad\quad\quad = \frac{7e^2}{48\pi^2}\frac{1}{\epsilon}[(k^2 g^{\lambda\nu} - k^\lambda k^\nu) - (q^2 g^{\nu\lambda} - q^\nu q^\lambda)].
\end{eqnarray}
Let us now look to the right-hand side of the identity~(\ref{ITF}). The diagrams are
\begin{eqnarray}
\raisebox{-0.4cm}{\incps{photonself1.eps}{-1.8cm}{-0.5cm}{2cm}{1.5cm}} &=& i\Pi^{\nu\lambda}_b(q) \\ \nonumber\\ \nonumber\\
&=& -i^2(2e)^2 \int \frac{d^4l}{(2\pi)^4} \frac{l^\lambda(l+q)^\nu}{l^2(q+l)^2} \sin^2(l\wedge q), \nonumber\\
\raisebox{-0.4cm}{\incps{photonself2.eps}{-1.8cm}{-0.5cm}{2cm}{1.5cm}} &=& i\Pi^{\nu\lambda}_c(q) \\ \nonumber\\ \nonumber\\
&=& \frac12(-i)^2(2e)^2 \int \frac{d^4l}{(2\pi)^4} \frac{\gamma^{\alpha\nu\beta}(l,q,-l-q)\gamma^{\beta\nu\alpha}(l+q,-q,-l)}{l^2(q+l)^2} \sin^2(l\wedge q), \nonumber\\
\raisebox{-1.2cm}{\incps{photonself3.eps}{-1.8cm}{-0.5cm}{1.5cm}{2.5cm}} &=& i\Pi^{\nu\lambda}_d(q) \\ \nonumber\\ \nonumber\\
&=& \frac12(-i)(-4ie^2) \int \frac{d^4l}{(2\pi)^4} \frac{{2(g^\alpha}_\alpha-1)g^{\nu\lambda}}{l^2} \sin^2(l\wedge q), \nonumber
\end{eqnarray}
so that
\begin{eqnarray}
{\rm PP}[\Pi^{\nu\lambda}_{b,c,d}(q) - \Pi^{\lambda\nu}_{b,c,d}(k)] &=& \frac{5e^2}{24\pi^2}\frac{1}{\epsilon}[(k^2 g^{\lambda\nu} - k^\lambda k^\nu) - (q^2 g^{\nu\lambda} - q^\nu q^\lambda)].
\end{eqnarray}
Finally, let us see the diagrams $B^\lambda_{\;\;\nu'}(p,q,k)$ and $B^\nu_{\;\;\lambda'}(p,k,q)$, given by
\begin{eqnarray}
\raisebox{-0.4cm}{\incps{Blambda1.eps}{-1.8cm}{-0.5cm}{1.2cm}{2cm}} &=&2e\sin(p\wedge q)B^\lambda_{a\,\nu'}(p,q,k) \\ \nonumber\\ \nonumber\\
&=& i^2(-i)(2e)^3 \int \frac{d^4l}{(2\pi)^4}\frac{(-l)_{\nu'}(-p-q-l)^\lambda}{l^2(p+l)^2(p+q+l)^2}\nonumber\\
&&\times\sin(l\wedge p)\sin(-l\wedge p - l\wedge q)\sin(-l\wedge q - p\wedge q), \nonumber\\
\raisebox{-0.4cm}{\incps{Blambda2.eps}{-1.8cm}{-0.5cm}{1.2cm}{2cm}} &=&2e\sin(p\wedge q)B^\lambda_{b\,\nu'}(p,q,k) \\ \nonumber\\ \nonumber\\
&=& i(-i)^2(2e)^3 \int \frac{d^4l}{(2\pi)^4}\frac{(p+l)_{\alpha}{\gamma^{\alpha\lambda}}_{\nu'}(-l,-p-q,p+q+l)}{l^2(p+l)^2(p+q+l)^2}\nonumber\\
&&\times\sin(l\wedge p)\sin(-l\wedge p - l\wedge q)\sin(l\wedge q + p\wedge q), \nonumber\\
\raisebox{-0.4cm}{\incps{Bnu1.eps}{-1.8cm}{-0.5cm}{1.2cm}{2cm}} &=&2e\sin(p\wedge q)B^\nu_{a\,\lambda'}(p,k,q) \\ \nonumber\\ \nonumber\\
&=& i^2(-i)(2e)^3 \int \frac{d^4l}{(2\pi)^4}\frac{(p+l)_{\lambda'}(p+q+l)^\nu}{l^2(p+l)^2(p+q+l)^2}\nonumber\\
&&\times\sin(l\wedge p)\sin(l\wedge p + l\wedge q)\sin(l\wedge q + p\wedge q),\quad\mathrm{and} \nonumber\\
\raisebox{-0.4cm}{\incps{Bnu2.eps}{-1.8cm}{-0.5cm}{1.2cm}{2cm}} &=&2e\sin(p\wedge q)B^\nu_{b\,\lambda'}(p,q,k) \\ \nonumber\\ \nonumber\\
&=& i(-i)^2(2e)^3 \int \frac{d^4l}{(2\pi)^4}\frac{(-l)_{\alpha}{\gamma^{\alpha\nu}}_{\lambda'}(p+l,q,-p-q-l)}{l^2(p+l)^2(p+q+l)^2}\nonumber\\
&&\times\sin(l\wedge p)\sin(-l\wedge p - l\wedge q)\sin(l\wedge q + p\wedge q), \nonumber
\end{eqnarray}
where their PP contributions are
\begin{eqnarray}
&& {\rm PP}[B^\lambda_{\;\;\nu'}(p,q,k)(q^2 g^{\nu\nu'} - q^\nu q^{\nu'}) + B^\nu_{\;\;\lambda'}(p,k,q)(k^2 g^{\lambda\lambda'} - k^\lambda k^{\lambda'})] \nonumber\\
&&\quad\quad\quad\quad\quad\quad = -\frac{e^2}{16\pi^2}\frac{1}{\epsilon}[(k^2 g^{\lambda\nu} - k^\lambda k^\nu) - (q^2 g^{\nu\lambda} - q^\nu q^\lambda)].
\end{eqnarray}
From this, we see that the renormalization constant $\tilde Z_1$ must be
\begin{equation}
\tilde Z_1= 1-\frac{e^2}{16\pi^2} \frac{1}{\epsilon},
\end{equation}
so that $\tilde Z_1=Z_5$ and thus we obtain the relations
\begin{equation}
Z_2/Z_{1F} = \tilde Z_3/\tilde Z_1 = Z_3/Z_1.
\end{equation}
Therefore, the sum of all PP contributions of the right-hand side of Eq.~(\ref{ITF}) becomes
\begin{eqnarray}
&& PP[\Pi^{\nu\lambda}_{b,c,d}(q) - \Pi^{\lambda\nu}_{b,c,d}(k)+{B^{\lambda}}_{\nu'}(p,q,k)(q^2 g^{\nu\nu'} - q^\nu q^{\nu'}) + {B^{\nu}}_{\lambda'}(p,q,k)(k^2 g^{\lambda\lambda'} - q^\lambda q^{\lambda'})] \nonumber\\
&&\quad\quad\quad\quad\quad\quad = \frac{7e^2}{48\pi^2}\frac{1}{\epsilon}[(k^2 g^{\lambda\nu} - k^\lambda k^\nu) - (q^2 g^{\nu\lambda} - q^\nu q^\lambda)],
\end{eqnarray}
which is the same result as for the left-hand side, Eq.~(\ref{sumL}).
Besides these ultraviolet divergent parts, arising from the planar parts of the diagrams, we have also infrared singular parts (SP) coming from the nonplanar parts of the same diagrams, at $p,q,k=0$. Explicit calculations, combining denominators with Feynman parameters and using nonplanar integrals, give us the SP for the diagrams $\Lambda^{\mu\nu\lambda}$ on the left-hand side of Eq.~(\ref{ITF}):
\begin{equation}
{\rm SP}[2e\sin(p\wedge q)\Lambda_{b1,b2,c,d1,d2,d3}^{\mu\nu\lambda}(p,q,k)] = \frac{4e^3}{\pi^2}\frac{\sin(p\wedge q)}{p\wedge q}\left(\frac{\tilde p^\mu\tilde p^\nu\tilde p^\lambda}{\xi\,\tilde p^4}+\frac{\tilde q^\mu\tilde q^\nu\tilde q^\lambda}{\xi\,\tilde q^4}+\frac{\tilde k^\mu\tilde k^\nu\tilde k^\lambda}{\xi\,\tilde k^4}\right),
\end{equation}
where we are not taking into account the logarithmic singularities. Contracting $q_\mu$ in the above expression, we obtain
\begin{equation}
{\rm SP}[q_\mu\Lambda_{b1,b2,c,d1,d2,d3}^{\mu\nu\lambda}(p,q,k)] = \frac{2e^2}{\pi^2}\left(\frac{\tilde q^\nu \tilde q^\lambda}{\xi^2\tilde q^4}-\frac{\tilde k^\nu \tilde k^\lambda}{\xi^2\tilde k^4}\right),
\end{equation}
which is exactly the same SP for the photon self-energy diagrams on the right-hand side of Eq.~(\ref{ITF}),
\begin{eqnarray}
&& {\rm SP}[\Pi^{\nu\lambda}_{b,c,d}(q) - \Pi^{\lambda\nu}_{b,c,d}(k)] = \frac{2e^2}{\pi^2}\left(\frac{\tilde q^\nu \tilde q^\lambda}{\xi^2\tilde q^4}-\frac{\tilde k^\nu \tilde k^\lambda}{\xi^2\tilde k^4}\right).
\end{eqnarray}
The other diagrams of the ST identity (\ref{ITF}) contribute only with logarithmic SP. These singularities are not problematic as they are integrable.
\section{Final comments}
In this work, for some specific Green functions, we have analysed the ST identities in the context of noncommutative QED$_4$. Special attention was given to the vectorial fermion-photon and triple photon vertex functions, explicitly verifying that no anomalies arise. The validity of these identities imply that, in spite of the presence of dangerous infrared singularities, the ultraviolet structure is not essentially modified by the noncommutativity. In fact, although the individual pole parts have been changed and new divergences appeared, the counterterms are related as they should in a non-Abelian situation. This however does not preclude the occurrence of dangerous infrared singularities which, in higher orders, jeopardizes the perturbative series. To extend our results to higher orders, our study must therefore be supplemented by some mechanism to control the mentioned singularities.
One possibility is to consider the effect of supersymmetry; as known supersymmetric theories have a better ultraviolet behavior and consequently
they may be free from dangerous infrared/ultraviolet mixing. This is what happens in susy noncommutative QED$_4$ \cite{ferrari}.
|
\section{Introduction}
For the past years, there have been great successes in measurement of CMB anisotropy by ground and satellite observations \citep{WMAP7:powerspectra,WMAP7:basic_result,WMAP5:basic_result,WMAP5:powerspectra,WMAP5:parameter,ACBAR,ACBAR2008,QUaD1,QUaD2,QUaD:instrument,Planck_bluebook}.
CMB anisotropy, which are associated with the inhomogeneity of the last scattering surface, provides the deepest survey so far and allows us to
constrain cosmological models significantly.
Since the initial release of WMAP data, the WMAP CMB sky data have undergone scrutiny, and various anomalies have been found and reported \citep{cold_spot1,cold_spot2,cold_spot_wmap3,cold_spot_origin,Tegmark:Alignment,Multipole_Vector1,Multipole_Vector2,Multipole_Vector3,Multipole_Vector4,Axis_Evil,Axis_Evil2,Axis_Evil3,Universe_odd,Park_Genus,Chiang_NG,Hemispherical_asymmetry,power_asymmetry_subdegree,power_asymmetry_wmap5,alfven,fnl_power,odd,lowl_anomalies}.
In particular, CMB anisotropy at low multipoles are associated with scales far beyond any existing astrophysical survey, and therefore
CMB anomalies at low multipoles may hint new physical laws at unexplored large scales.
CMB sky map may be considered as the sum of even and odd parity functions. Previously, Land and et al. have noted the odd point-parity preference of WMAP CMB data, but found its statistical significance is not high enough \citep{Universe_odd}.
In our previous work, we have applied a slight different estimator to WMAP 5 year power spectrum data, and found
significant odd point-parity preference at low multipoles \citep{odd}.
In this paper, we investigate the recently released WMAP 7 year data up to higher multipoles, and discuss origins of the observed odd-parity preference.
This paper is organized as follows.
In Section \ref{asymmetry}, we discuss the anomalous odd-parity preference of the WMAP data.
In Section \ref{cosmomc}, we implement cosmological fitting by excluding even or odd low multipole data.
In Section \ref{systematics} and \ref{cosmic}, we investigate non-cosmological and cosmological origin.
In Section \ref{discussion}, we summarize our investigation and discuss prospect.
In Appendix \ref{CMB}, we briefly review statistical properties of Gaussian CMB anisotropy.
\section{Parity asymmetry of the WMAP data}
\label{asymmetry}
CMB anisotropy sky map may be considered as the sum of even and odd parity functions:
\begin{eqnarray}
T(\hat{\mathbf n})=T^+(\hat{\mathbf n})+T^-(\hat{\mathbf n}),
\end{eqnarray}
where
\begin{eqnarray}
T^+(\hat{\mathbf n})&=&\frac{T(\hat{\mathbf n})+T(-\hat{\mathbf n})}{2},\\
T^-(\hat{\mathbf n})&=&\frac{T(\hat{\mathbf n})-T(-\hat{\mathbf n})}{2}.
\end{eqnarray}
Taking into account the parity property of spherical harmonics $Y_{lm}(\hat{\mathbf n})=(-1)^l\,Y_{lm}(-\hat{\mathbf n})$ \citep{Arfken},
we may easily show
\begin{eqnarray}
T^+(\hat{\mathbf n})&=&\sum_{l=2n,m} a_{lm}\,Y_{lm}(\hat{\mathbf n}), \label{T_even}\\
T^-(\hat{\mathbf n})&=&\sum_{l=2n-1,m} a_{lm}\,Y_{lm}(\hat{\mathbf n}), \label{T_odd}
\end{eqnarray}
where $n$ is an integer.
Therefore, significant power asymmetry between even and odd multipoles indicates a preference for a particular parity.
Hereafter, we will denote a preference for particular parity by `parity asymmetry'.
In Fig. \ref{Cl_cut}, we show the WMAP 7 year, 5 year, 3 year data and the WMAP concordance model \citep{WMAP7:powerspectra,WMAP5:powerspectra,WMAP3:temperature,WMAP5:Cosmology,WMAP7:Cosmology}.
\begin{figure}[htb!]
\centering\includegraphics[scale=.5]{./figure/fig1.eps}
\caption{CMB power spectrum: WMAP 7 year data (blue), WMAP 5 year data (green) and WMAP 3 year data (red), $\Lambda$CDM model (cyan)}
\label{Cl_cut}
\end{figure}
We may see from Fig. \ref{Cl_cut} that the power spectrum of WMAP data at even multipoles tend to be lower than those at neighboring odd multipoles.
\begin{figure}[htb]
\centering\includegraphics[scale=.5]{./figure/fig2.eps}
\caption{$(-1)^l\times$ difference between WMAP power spectrum data and $\Lambda$CDM model}
\label{delta}
\end{figure}
In Fig. \ref{delta}, we show $(-1)^l l(l+1)/2\pi\:(C^{\mathrm{WMAP}}_l-C^{\Lambda\mathrm{CDM}}_l)$ for low multipoles.
Since we expect random scattering of data points around a theoretical model, we expect the distribution of dots in Fig. \ref{delta} to be symmetric around zero.
However, there are only 5 points of positive values among 22 points in the case of WMAP7 or WMAP5 data.
Therefore, we may see that there is the tendency of power deficit (excess) at even (odd) multipoles, compared with the $\Lambda$CDM model.
Taking into account $l(l+1) C_l\sim \mathrm{const}$, we may consider the following quantities:
\begin{eqnarray}
P^{+} &=& \sum^{l_{\mathrm{max}}}_{l=2} 2^{-1}\left(1+(-1)^{l}\right)\, l(l+1)/2\pi \: C_l,\\
P^{-} &=& \sum^{l_{\mathrm{max}}}_{l=2} 2^{-1}\left(1-(-1)^{l}\right)\, l(l+1)/2\pi \: C_l,
\end{eqnarray}
where $P^{+}$ and $P^{-}$ are the sum of $l(l+1)/2\pi \: C_l$ for even and odd multipoles respectively.
Therefore, the ratio $P^{+}/P^{-}$ is associated with the degree of the parity asymmetry, where the lower value of $P^+/P^-$ indicates odd-parity preference, and vice versa. It is worth to note that our estimator of the parity asymmetry does not possess explicit dependence on the underlying theoretical model in the sense that the term $C^{\mathrm{WMAP}}_l-C^{\Lambda\mathrm{CDM}}_l$ is absent in our estimator.
\begin{figure}[htb]
\centering\includegraphics[scale=.5]{./figure/fig3.eps}
\caption{$P^+/P^-$ of WMAP data and $\Lambda$CDM}
\label{P_ratio}
\end{figure}
In Fig. \ref{P_ratio}, we show the $P^+/P^-$ of WMAP data, and a $\Lambda$CDM model for various $l_{\mathrm{max}}$.
As shown in Fig. \ref{P_ratio}, $P^+/P^-$ of WMAP data are far below theoretical values. Though the discrepancy is largest at lowest $l_{\mathrm{max}}$, its statistical significance is not necessarily high for low $l$, due to large statistical fluctuation.
In order to make rigorous assessment on its statistical significance at low $l$, we are going to compare $P^+/P^-$ of WMAP data with simulation.
We have produced $10^4$ simulated CMB maps HEALPix Nside=8 and Nside=512 respectively, via map synthesis with $a_{lm}$ randomly drawn from Gaussian $\Lambda$CDM model.
We have degraded the WMAP processing mask (Nside=16) to Nside=8, and set pixels to zero, if any of their daughter pixels is zero.
After applying the mask, we have estimated power spectrum $2\le l\le 23$ from simulated cut-sky maps (Nside=8) by a pixel-based maximum likelihood method \citep{WMAP7:powerspectra,Bond:likelihood,hybrid_estimation}.
At the same time, we have applied the WMAP team's KQ85 mask to the simulated maps (Nside=512), and estimated power spectrum $2\le l\le 1024$ by pseudo $C_l$ method \citep{pseudo_Cl,MASTER}. In the simulation, we have neglected instrument noise, since the signal-to-noise ratio of the WMAP data is quite high at multipoles of interest (i.e. $l\le 100$) \citep{WMAP7:powerspectra,WMAP7:basic_result}. Using the low $l$ estimation by pixel-maximum likelihood method and high $l$ estimation by pseudo $C_l$ method, we have computed $P^+/P^-$ respectively for various multipole ranges $2\le l \le l_{\mathrm{max}}$, and compared $P^+/P^-$ of the WMAP data with simulation.
\begin{figure}[htb]
\centering\includegraphics[scale=.5]{./figure/fig4.eps}
\caption{Probability of getting $P^+/P^-$ as low as WMAP data for multipole range $2\le l\le{l_\mathrm{max}}$.}
\label{P}
\end{figure}
In Fig. \ref{P}, we show $p$-value of WMAP7, WMAP5 and WMAP3 respectively for various $l_{\mathrm{max}}$, where $p$-value denotes fractions of simulations as low as $P^+/P^-$ of the WMAP data. As shown in Fig. \ref{P}, the parity asymmetry of WMAP7 data at multipoles ($2\le l\le 22$) is most anomalous, where $p$-value is $0.0031$.
As shown in Fig. \ref{P}, the statistical significance of the parity asymmetry (i.e. low $p$-value) is getting higher, when we include higher multipoles up to 22.
Therefore, we may not attribute the odd parity preference simply to the low quadrupole power, and find it rather likely that the low quadrupole power is not an isolated anomaly, but shares an origin with the odd parity preference.
\begin{table}[htb]
\centering
\caption{the parity asymmetry of WMAP data ($2\le l\le 22$)}
\begin{tabular}{ccc}
\hline\hline
data & $P^+/P^-$ & $p$-value\\
\hline
WMAP7 & 0.7076 & 0.0031\\
WMAP5 & 0.7174 & 0.0039\\
WMAP3 & 0.7426 & 0.0061\\
\hline
\end{tabular}
\label{pvalue}
\end{table}
In Table \ref{pvalue}, we summarize $P^+/P^-$ and $p$-values of WMAP7, WMAP5 and WMAP3 for $l_{\mathrm{max}}=22$. As shown in Fig. \ref{P} and Table \ref{pvalue}, the odd-parity preference of WMAP7 is most anomalous, while WMAP7 data are believed to have more accurate calibration and less foreground contamination than earlier releases. \citep{WMAP7:powerspectra,WMAP7:basic_result,WMAP5:basic_result,WMAP5:powerspectra,WMAP5:beam}.
In Fig. \ref{hist}, we show cumulative distribution of $P^+/P^-$ for $10^4$ simulated maps. The values corresponding to $P^+/P^-$ of WMAP data are marked as dots.
\begin{figure}[htb]
\centering\includegraphics[scale=.5]{./figure/fig5.eps}
\caption{Parity asymmetry at multipoles ($2\le l\le 22$): cumulative distribution of $P^+/P^-$ for $10^4$ simulated maps (cyan), $P^+/P^-$ of WMAP7 (blue), WMAP5 (green) and WMAP3 (red)}
\label{hist}
\end{figure}
We have also compared $P^+/P^-$ of the WMAP7 with whole-sky simulation (i.e. no mask), and obtained $p$-value $0.002$ for $l_{\mathrm{max}}=22$.
The lower $p$-value from whole-sky simulation is attributed to the fact that statistical fluctuation in whole-sky $C_l$ estimation is smaller than that of cut-sky estimation.
In the absence of strong theoretical grounds for the parity asymmetry ($2\le l\le 22$),
we have to take into account our posteriori choice on $l_{\mathrm{max}}$, which might have enhanced the statistical significance.
In order to do that, we have produced whole-sky Monte-Carlo simulations, and retained only simulations, whose $p$-value is lowest at $l_{\mathrm{max}}=22$.
The $p$-values have been estimated by comparing them with simulations, which are produced separately.
Once we have retained $10^4$ simulations, we have compared them with WMAP7 data, and found that only fraction $0.0197$ of retained simulations have $P^+/P^-$ as low as WMAP7 data.
The statistical significance of the parity asymmetry ($2\le l\le 22$) is reduced substantially by accounting for the posteriori choice on $l_{\mathrm{max}}$.
However, it is still significant.
We have also investigated the parity asymmetry with respect to mirror reflection (i.e. $l+m$=even or odd) respectively in Galactic coordinate and Ecliptic coordinate. However, we find the statistical significance is not as high as the point-parity asymmetry (i.e. $l$=even or odd).
\section{the power contrast and the $\Lambda$CDM model fitting}
\label{cosmomc}
The parity asymmetry discussed in the previous section is explicitly associated with the angular power spectrum data, which are used extensively to fit cosmological models.
Noting the significant power contrast between even and odd low multipoles, we have investigated cosmological models respectively by excluding even or odd low multipole data ($2\le l \le 22$) from total dataset.
Hereafter, we denote CMB data of even(odd) multipoles ($2\le l\le 22$) plus all high multipoles as `$D_2$'(`$D_3$') respectively.
For a cosmological model, we have considered $\Lambda$CDM + SZ effect + weak-lensing, where cosmological parameters are $\lambda \in \left\{\Omega_b,\Omega_{c},\tau,n_s, A_s, A_{sz}, H_0 \right\}$. For data constraints, we have used the WMAP 7 year power spectrum data \citep{WMAP7:powerspectra}.
In order to exclude even or odd low multipoles, we have made slight modifications to the likelihood code provided by the WMAP team, and run a \texttt{CosmoMC} with the modification. \citep{Gibbs_power,WMAP7:powerspectra,CosmoMC}.
\begin{figure}[htb]
\centering
\includegraphics[scale=.54]{./figure/fig6.eps}
\caption{Marginalized likelihood of cosmological parameters: results are obtained respectively with/without even or odd multipole data ($2\le l \le 22$).}
\label{like}
\end{figure}
In Fig. \ref{like}, we show the marginalized likelihoods of parameters.
As shown in Fig. \ref{like}, the parameter likelihood imposed by $D_3$ seem to differ from those of others.
In Table \ref{parameter}, we show the bestfit parameters and 1 $\sigma$ confidence intervals, where $\lambda$, $\lambda_2$ and $\lambda_3$ denote the bestfit values of the full data, $D_2$ and $D_3$ respectively.
\begin{table}[htb]
\centering
\caption{cosmological parameters ($\Lambda$CDM + sz + lens)}
\begin{tabular}{cccc}
\hline\hline
& $\lambda$ &$\lambda_2$ & $\lambda_3$ \\
\hline
$\Omega_{b}\,h^2$ & $0.0226\pm 0.0006$ &$0.0228\pm0.0006$ & $0.0221\pm0.0006$ \\
$\Omega_{c}\,h^2$ & $0.112\pm0.006$ &$0.11\pm0.006$ & $0.116\pm0.006$ \\
$\tau$ & $0.0837\pm 0.0147$ &$0.0879\pm0.015$ & $0.087\pm0.0147$ \\
$n_s$ & $0.964\pm 0.014$ &$0.974\pm0.015$ & $0.95\pm0.015$ \\
$\log[10^{10} A_s]$ & $3.185\pm0.047$ &$ 3.165\pm0.049$ & $3.246\pm0.048$ \\
$H_0$ & $70.53\pm2.48$ &$71.43\pm2.51$ & $68.07\pm2.53$ \\
$A_{\mathrm{sz}}$ & $1.891^{+0.109}_{-1.891}$ &$1.469^{+0.541}_{-1.469}$ & $1.558^{+0.442}_{-1.558}$ \\
\hline
\end{tabular}
\label{parameter}
\end{table}
As mentioned above, there exist some level of tension between $D_3$ and the full data.
However, the Bayes Factor $\mathcal{L}(\lambda|D_3)/\mathcal{L}(\lambda_3|D_3)= \exp(-3733.124+3732.791)=0.71$ shows that
it is `not worth than a bare mention', according to Jeffreys' scale \citep{Jeffreys_scale}.
\section{Non-cosmological origins}
\label{systematics}
In this section, we are going to investigate non-cosmological origins such as asymmetric beams, instrument noise, foreground and cut-sky effect.
\subsection{asymmetric beam}
\begin{figure}[htb]
\centering
\includegraphics[scale=.25]{./figure/fig7a.eps}
\includegraphics[scale=.25]{./figure/fig7b.eps}
\includegraphics[scale=.25]{./figure/fig7c.eps}
\includegraphics[scale=.25]{./figure/fig7d.eps}
\includegraphics[scale=.25]{./figure/fig7e.eps}
\includegraphics[scale=.25]{./figure/fig7f.eps}
\caption{the parity asymmetry in the presence of beam asymmetry:
The dots denotes ($P^+$,$P^-$) of CMB maps simulated with asymmetric beams.
The dashed lines are plotted with slopes corresponding to the $P^+/P^-$ of $\Lambda$CDM model (red) and WMAP7 data (green) respectively.
The alphanumeric values at the lower right corner denote the frequency band and D/A channel.}
\label{beam}
\end{figure}
The shape of the WMAP beams are slightly asymmetric \citep{WMAP5:beam,WMAP3:beam,asymmetric_beam}, while
the WMAP team have assumed symmetric beams in the power spectrum estimation \citep{WMAP7:powerspectra,WMAP5:powerspectra,WMAP5:beam,WMAP3:beam}.
We have investigated the association of beam asymmetry with the anomaly, by relying on
simulated maps provided by \citep{asymmetric_beam}.
The authors have produced 10 simulated maps for each frequency and Differencing Assembly (D/A) channels, where the detailed shape of the WMAP beams and the WMAP scanning strategy
are taken into account \citep{asymmetric_beam}.
From simulated maps, we have estimated $P^+$ and $P^-$, where we have compensated for beam smoothing purposely by the WMAP team's beam transfer function (i.e. symmetric beams).
In Fig. \ref{beam}, we show $P^+$ and $P^-$ values of the simulated maps, and the dashed lines of a slope corresponding to $P^+/P^-$ of $\Lambda$CDM and WMAP7 data respectively.
As shown in Fig. \ref{beam}, we do not observe the odd-parity preference of WMAP data in simulated maps.
Therefore, we find it hard to attribute the odd-parity preference to asymmetric beams.
\subsection{noise}
There exist instrument noise in the WMAP data.
Especially, 1/f noise, when coupled with WMAP scanning pattern, may result in less accurate measurement at certain low multipoles \citep{WMAP3:temperature,Detection_Light,WMAP1:processing}.
In order to investigate the association of noise with the anomaly, we have produced noise maps of WMAP7 data by subtracting one Differencing Assembly (D/A) map from another D/A data of the same frequency channel.
In Fig. \ref{noise}, we show $P^+$ and $P^-$ values of the noise maps.
\begin{figure}[htb]
\centering
\includegraphics[scale=.5]{./figure/fig8.eps}
\caption{the parity asymmetry of the WMAP noise:
The dots denotes ($P^+$,$P^-$) of noise maps, and alphanumeric values in the legend denote the frequency band and the pair of D/A channels used. Two dashed lines are plotted with the slope corresponding to $P^+/P^-$ of white noise and WMAP7 data respectively.}
\label{noise}
\end{figure}
As shown in Fig. \ref{noise}, the noise maps do not show odd-parity preference, but their $P^+/P^-$ ratios are consistent with that of white noise (i.e. $C_l=\mathrm{const.}$).
Besides that, the signal-to-noise ratio of WMAP temperature data is quite high at low multipoles (e.g S/N$\sim$ 100 for $l=30$) \citep{WMAP3:temperature,WMAP5:beam,WMAP1:processing}.
Therefore, we find that instrument noise, including 1/f noise, is unlikely to be the cause of the odd-parity preference.
\subsection{foreground}
There are contamination from galactic and extragalactic foregrounds.
In order to reduce foreground contamination, the WMAP team have subtracted diffuse foregrounds by template-fitting, and masked the regions that cannot be cleaned reliably.
For foreground templates (dust, free-free emission and synchrotron), the WMAP team used dust emission ``Model 8", H$\alpha$ map, and the difference between K and Ka band maps \citep{WMAP5:basic_result,WMAP7:fg,Dust_Extrapolation,Finkbeiner_H_alpha,WMAP1:fg}.
In Fig. \ref{template}, we show the power spectrum of templates.
\begin{figure}[htb]
\centering
\includegraphics[scale=.5]{./figure/fig9.eps}
\caption{the power spectra of the templates (synchrotron, H$\alpha$, dust): plotted with arbitrary normalization.}
\label{template}
\end{figure}
As shown in Fig. \ref{template}, templates show strong even parity preference, which is opposite to that of the WMAP power spectrum data.
Therefore, one might attribute the odd-parity preference of WMAP data to over-subtraction by templates.
However, we find the error associated with template-fitting is unlikely to be the cause of the odd-parity preference.
Consider spherical harmonic coefficients of a foreground-reduced map:
\begin{eqnarray}
a^{\mathrm{obs}}_{lm}=a^{\mathrm{cmb}}_{lm}+a^{\mathrm{fg}}_{lm}-b\,a^{\mathrm{tpl}}_{lm},
\end{eqnarray}
where $a^{\mathrm{obs}}_{lm}$, $a^{\mathrm{fg}}_{lm}$ and $b\,a^{\mathrm{tpl}}_{lm}$ correspond to a foreground-cleaned map, a foreground and
a template with a fitting coefficient $b$.
For simplicity, we consider only a single foreground component, but the conclusion is equally valid for multi-component foregrounds.
Since there is no correlation between foregrounds and CMB, the observed power spectrum is given by:
\begin{eqnarray}
C^{\mathrm{obs}}_l\approx C^{\mathrm{cmb}}_l+ \langle \left|a^{\mathrm{fg}}_{lm} -b\, a^{\mathrm{tpl}}_{lm} \right|^2\rangle. \label{Cl_obs}
\end{eqnarray}
As shown Eq. \ref{Cl_obs}, the parity preference should follow that of templates because of the second term, provided templates are good tracers of foregrounds (i.e. $a^{\mathrm{fg}}_{lm}/a^{\mathrm{tpl}}_{lm}\approx \mathrm {const}$). Nevertheless, Eq. \ref{Cl_obs} may make a bad approximation for lowest multipoles, because the cross term $\sum_m \mathrm{Re}[a^{\mathrm{cmb}}_{lm} (a^{\mathrm{fg}}_{lm} -b\, a^{\mathrm{tpl}}_{lm})^*]$ may not be negligible.
Besides that, our argument and the template-fitting method itself fail, if templates are not good tracers of foregrounds.
In order to investigate these issues, we have resorted to simulation in combination with WMAP data.
Noting the WMAP power spectrum is estimated from foreground-reduced V and W band maps, we have produced simulated maps as follows:
\begin{eqnarray}
T(\hat{\mathbf n})=T_{\mathrm{cmb}}(\hat{\mathbf n})+(V(\hat{\mathbf n})-W(\hat{\mathbf n}))/2, \label{residual_fg2}
\end{eqnarray}
where $V(\hat{\mathbf n})$ and $W(\hat{\mathbf n})$ are foreground-reduced V and W band maps of WMAP data. Note that the second term on the right hand side contains only residual foregrounds at V and W band, because the difference of distinct frequency channels are mainly residual foregrounds at low $l$.
Just as cut-sky simulation described in Sec. \ref{asymmetry}, we have applied a foreground mask to the simulated maps, and estimated the power spectrum from cut-sky by a pixel-based maximum likelihood method.
\begin{figure}[htb]
\centering
\includegraphics[scale=.5]{./figure/fig10.eps}
\caption{the parity asymmetry in the presence of residual foregrounds (V$-$W): Dashed lines are plotted with slopes corresponding to $P^+/P^-$ of $P^+/P^-$ of $\Lambda$CDM (cyan), WMAP7 data (red).}
\label{residual2}
\end{figure}
In Fig. \ref{residual2}, we show $P^+$ and $P^-$ values estimated from simulations.
For comparison, we have included simulations without residual foregrounds, and dashed lines of a slope corresponding to $P^+/P^-$ of $\Lambda$CDM model and WMAP7 data. As shown in Fig. \ref{residual2}, the $P^+/P^-$ of simulations in the presence of residual foregrounds do not show anomalous odd-parity preference of WMAP data.
Considering Eq. \ref{Cl_obs} and simulations, we find it difficult to attribute the odd-parity preference to residual foreground.
There also exist contamination from unresolved extragalactic point sources \citep{WMAP3:temperature}.
However, point sources follow Poisson distribution with little departure \citep{Tegmark:Foreground}, and therefore are unlikely to possess odd-parity preference.
Besides that, point sources at WMAP frequencies are subdominant on large angular scales (low $l$) \citep{WMAP3:temperature,WMAP7:fg,Tegmark:Foreground,WMAP5:fg}.
Though we have not find association of foregrounds with the anomaly,
we do not rule out residual foreground completely, because of our limited knowledge on residual foregrounds.
\subsection{cut sky}
The WMAP team have masked the region that cannot be reliably cleaned by template fitting, and estimated CMB power spectrum from sky data outside the mask \citep{WMAP7:powerspectra,WMAP5:powerspectra,WMAP3:temperature,WMAP7:fg}.
Therefore, we have estimated the $p$-value presented in Sec. \ref{asymmetry}, by comparing WMAP data with cut-sky simulations.
\begin{figure}[htb]
\centering
\includegraphics[scale=.5]{./figure/fig11.eps}
\caption{Probability of getting $P^+/P^-$ as low as the ILC 7 year, 5 year, and 3 year map at multipole range $2\le l\le{l_\mathrm{max}}$}
\label{P_wilc}
\end{figure}
Nonetheless, we have investigated the WMAP team's Internal Linear Combination map (ILC) map in order to see if the odd-parity preference also exists in a whole-sky CMB map.
Note that the WMAP ILC map provides a reliable estimate of CMB signal over whole-sky on angular scales larger than $10^\circ$ \citep{WMAP3:temperature,WMAP7:fg,WMAP5:fg}.
We have compared $P^+/P^-$ of the ILC maps with whole-sky simulations.
In Fig. \ref{P_wilc}, we show $p$-values of the ILC maps respectively for various $l_{\mathrm{max}}$.
As shown in Fig. \ref{P_wilc}, the odd-parity preference of ILC maps is most anomalous for $l_{\mathrm{max}}=22$ as well.
In Table \ref{pvalue_wilc}, we summarize $P^+/P^-$ and $p$-values for $l_{\mathrm{max}}=22$.
\begin{table}[htb]
\centering
\caption{the parity asymmetry of WMAP ILC maps ($2\le l\le 22$)}
\begin{tabular}{ccc}
\hline\hline
data & $P^+/P^-$ & $p$-value\\
\hline
ILC7 & 0.7726 & 0.0088\\
ILC5 & 0.7673 & 0.0074\\
ILC3 & 0.7662 & 0.0072\\
\hline
\end{tabular}
\label{pvalue_wilc}
\end{table}
As shown in Fig. \ref{P_wilc} and Table \ref{pvalue_wilc}, we find anomalous odd-parity preference exits in whole-sky CMB maps as well.
Therefore, we find it difficult to attribute the anomaly to cut-sky effect.
\subsection{Other known sources of errors}
Besides contamination discussed in previous sections, there are other sources of contamination such as sidelobe pickup and so on.
In order to investigate these effects, we have resorted to simulation produced by the WMAP team.
According to the WMAP team, time-ordered data (TOD) have been simulated with realistic noise, thermal drifts in instrument gains and baselines,
smearing of the sky signal due to finite integration time, transmission imbalance, and far-sidelobe beam pickup.
Using the same data pipeline used for real data, the WMAP team have processed simulated TOD, and produced maps for each differencing assembly and each single year observation year.
\begin{figure}[htb]
\centering
\includegraphics[scale=.5]{./figure/fig12.eps}
\caption{ $P^+$ and $P^-$ of the WMAP team's simulation for V and W band data}
\label{sim}
\end{figure}
In Fig. \ref{sim}, we show the $P^+$ and $P^-$ of the simulated maps, where the power spectrum estimation is made from cut-sky by a pixel-based likelihood method.
As shown in Fig. \ref{sim}, all points are well above $P^+/P^-$ of WMAP7, and agree with $\Lambda$CDM model.
Therefore, we do not find definite association of the parity asymmetry with known systematics effects.
\section{Cosmological origin}
\label{cosmic}
In this section, we are going to take the WMAP power spectrum at face values, and consider cosmological origins.
Topological models including multi-connected Universe and Bianchi $VII$ model have been proposed to explain the cold spot or low quadrupole power \citep{low_quadrupole,spherical_tessellation,Land_Bianchi}.
However, the topological models do not produce the parity asymmetry, though some of them, indeed, predict low quadrupole power.
Trans-Planckian effects and some inflation models predict oscillatory features in primordial power spectrum \citep{Inflation,Inflation_Planckian_problem,Inflation_Planckian_spectra,Inflation_Planckian_note,Inflation_Planckian_estimate,CMB_Planckian_signature,WMAP_oscillation,Inflation_Planckian,Inflation_initial,Planckian_Astrophysics,CMB_Planckian_observation,WMAP3:parameter}.
However, oscillatory or sharp features in primordial power spectrum are smeared out in translation to the CMB power spectrum \citep{WMAP7:anomaly}.
Besides, reconstruction of primordial power spectrum and investigation on features show that primordial power spectrum is close to a featureless power-law spectrum \citep{WMAP7:powerspectra,WMAP5:Cosmology,WMAP7:Cosmology,WMAP3:parameter,power_recon,power_svd,power_features}.
Therefore, we find it difficult to attribute the anomaly to trans-Planckian effect or extended inflation models.
We will consider what the odd-parity preference imply on primordial perturbation $\Phi(\mathbf k)$, if primordial power spectrum is, indeed, featureless. Using Eq. \ref{alm}, we may show the decomposition coefficients of CMB anisotropy are given by:
\begin{eqnarray*}
a_{lm}&=&\frac{(-\imath)^l}{2\pi^2} \int\limits^{\infty}_0 dk \int\limits^{\pi}_0 d\theta_{\mathbf k} \sin\theta_{\mathbf k} \int\limits^{2\pi}_0 d\phi_{\mathbf k}\,\Phi(\mathbf k)\,g_{l}(k)\,Y^*_{lm}(\hat{\mathbf k}),\nonumber\\
&=&\frac{(-\imath)^l}{2\pi^2} \int\limits^{\infty}_0 dk \int\limits^{\pi}_0 d\theta_{\mathbf k} \sin\theta_{\mathbf k}\int\limits^{\pi}_0 d\phi_{\mathbf k}\,g_{l}(k)\times\\
&&\left(\Phi(\mathbf k)\,Y^*_{lm}(\hat {\mathbf k}) + \Phi(-\mathbf k)\,Y^*_{lm}(-\hat {\mathbf k}) \right),\nonumber\\
&=&\frac{(-\imath)^l}{2\pi^2} \int\limits^{\infty}_0 dk \int\limits^{\pi}_0 d\theta_{\mathbf k} \sin\theta_{\mathbf k}\int\limits^{\pi}_0 d\phi_{\mathbf k}\,g_{l}(k) Y^*_{lm}(\hat {\mathbf k})\times\\
&&\left(\Phi(\mathbf k)+(-1)^l \Phi^*(\mathbf k)\right),
\end{eqnarray*}
where we used the reality condition $\Phi(-\mathbf k)= \Phi^*(\mathbf k)$ and $Y_{lm}(\hat{-\mathbf n})=(-1)^l\,Y_{lm}(\hat{\mathbf n})$.
Using Eq. \ref{alm2}, it is trivial to show, for the odd number multipoles $l=2n-1$,
\begin{eqnarray}
\lefteqn{a_{lm}=}\label{alm2}\\
&&-\frac{(-\imath)^{l-1}}{\pi^2} \int\limits^{\infty}_0 dk \int\limits^{\pi}_0 d\theta_{\mathbf k} \sin\theta_{\mathbf k}\int\limits^{\pi}_0 d\phi_{\mathbf k}\,g_{l}(k) Y^*_{lm}(\hat {\mathbf k})\,\mathrm{Im}[\Phi(\mathbf k)],\nonumber
\end{eqnarray}
and, for even number multipoles $l=2n$,
\begin{eqnarray}
\lefteqn{a_{lm}=}\label{alm3}\\
&&\frac{(-\imath)^l}{\pi^2} \int\limits^{\infty}_0 dk \int\limits^{\pi}_0 d\theta_{\mathbf k} \sin\theta_{\mathbf k}\int\limits^{\pi}_0 d\phi_{\mathbf k}\,g_{l}(k) Y^*_{lm}(\hat {\mathbf k})\,\mathrm{Re}[\Phi(\mathbf k)]\nonumber.
\end{eqnarray}
It should be noted that the above equations are simple reformulation of Eq. \ref{alm}, and exactly equal to them.
From Eq. \ref{alm2} and \ref{alm3}, we may see that the odd-parity preference might be produced, provided
\begin{eqnarray}
|\mathrm{Re} [\Phi(\mathbf k)]|\ll|\mathrm{Im} [\Phi(\mathbf k)]|\;\;\;(k\lesssim 22/\eta_0),\label{primordial_odd}
\end{eqnarray}
where $\eta_0$ is the present conformal time.
Taking into account the reality condition $\Phi(-\mathbf k)= \Phi^*(\mathbf k)$, we may show primordial perturbation in real space is given by:
\begin{eqnarray}
\Phi(\mathbf x)&=&2\int\limits^{\infty}_0 dk \int\limits^{\pi}_0 d\theta_{\mathbf k} \sin\theta_{\mathbf k}\int\limits^{\pi}_0 d\phi_{\mathbf k}\label{Phi_real}\\
&\times&\left(\mathrm{Re}[\Phi(\mathbf k)]\cos(\mathbf k\cdot \mathbf x)-\mathrm{Im}[\Phi(\mathbf k)]\sin(\mathbf k\cdot \mathbf x)\right). \nonumber
\end{eqnarray}
Noting Eq. \ref{primordial_odd} and \ref{Phi_real}, we find our primordial Universe may possess odd-parity preference on large scales ($2/\eta_0 \lesssim k\lesssim 22/\eta_0$).
The odd-parity preference of our primordial Universe violates large-scale translational invariance in all directions.
However, it is not in direct conflict with the current data on observable Universe (i.e. WMAP CMB data), though it may seem intriguing.
Considering Eq. \ref{primordial_odd} and \ref{Phi_real}, we find this effect will be manifested on the scales larger than $2\pi\,\eta_0/22\approx 4\,\mathrm{Gpc}$.
However, it will be difficult to observe such large-scale effects in non-CMB observations.
If the odd-parity preference is indeed cosmological, it indicates we are at a special place in the Universe, which may sound bizarre.
However, it should be noted that the invalidity of the Copernican Principle such as our living near the center of void had been previously proposed in different context \citep{Void_DE,Void_SN}.
Depending on the type of cosmological origins (e.g. topology, features in primordial power spectrum and Eq. \ref{primordial_odd}), distinct anomalies are predicted in polarization power spectrum. Therefore, polarization maps of large-sky coverage (i.e. low multipoles) will allow us to remove degeneracy and
figure a cosmological origin, if the parity asymmetry is indeed cosmological.
\section{Discussion}
\label{discussion}
We have investigated the parity asymmetry of our early Universe, using the newly released WMAP 7 year power spectrum data.
Our investigation shows anomalous odd-parity preference of the WMAP7 data ($2\le l\le22$) at 3-in-1000 level.
When we account for our posteriori choice on $l_\mathrm{max}$, the statistical significance decreases to 2-in-100 level, but remains significant.
There exist several known CMB anomalies at low multipole, including the parity asymmetry discussed in this paper \citep{Tegmark:Alignment,Multipole_Vector1,Multipole_Vector2,Multipole_Vector3,Multipole_Vector4,Axis_Evil,Axis_Evil2,Axis_Evil3,odd,lowl,lowl_WMAP13} (see \citep{lowl_anomalies} for review). We find it likely there exist a common underlying origin, whether cosmological or not.
We have investigated non-cosmological origins, and ruled out various non-cosmological origins such as asymmetric beams, noise and cut-sky effect.
We have also investigated the WMAP team's simulation, which includes all known systematic effects, and do not find definite association with known systematics.
Among cosmological origins, topological models or primordial power spectrum of feature might provide theoretical explanation, though currently available models do not.
We also find primordial origin requires $|\mathrm{Re} [\Phi(\mathbf k)]|\ll|\mathrm{Im} [\Phi(\mathbf k)]|$ for $k\lesssim 22/\eta_0$, if we consider a simple phenomenologically fitting model. In other words, it requires violation of translation invariance in primordial Universe on the scales larger than $4\,\mathrm{Gpc}$.
Depending on the type of cosmological origins, distinct anomalies are predicted in polarization power spectrum.
Therefore, we will be able to remove degeneracy in cosmological origins, when polarization data of large sky coverage are available.
However, at this moment, it is not even clear, whether the anomaly is due to unaccounted contamination or indeed cosmological.
Nonetheless, we may be able to resolve the mystery of the large-scale odd-parity preference, when data from the Planck surveyor are available.
\section{Acknowledgments}
We are grateful to the anonymous referee for thorough reading and helpful comments, which leads to significant improvement of this work.
We thank Eiichiro Komatsu, Paolo Natoli and Dominik Schwarz for useful discussion.
We acknowledge the use of the Legacy Archive for Microwave Background Data Analysis (LAMBDA).
We acknowledge the use of the simulated CMB maps of asymmetric beams produced by Wehus et al. \citep{asymmetric_beam}.
Our data analysis made the use of HEALPix \citep{HEALPix:Primer,HEALPix:framework}.
This work is supported in part by Danmarks Grundforskningsfond, which allowed the establishment of the Danish Discovery Center.
This work is supported by FNU grant 272-06-0417, 272-07-0528 and 21-04-0355.
\begin{appendix}
\section{Statistical Properties of CMB anisotropy}
\label{CMB}
The CMB temperature anisotropy over a whole-sky is conveniently decomposed in terms of spherical harmonics $Y_{lm}(\theta,\phi)$ as follows:
\begin{eqnarray}
T(\hat{\mathbf n})=\sum_{lm} a_{lm}\,Y_{lm}(\hat{\mathbf n}),
\end{eqnarray}
where $a_{lm}$ is a decomposition coefficient, and $\hat{\mathbf n}$ is a sky direction.
Decomposition coefficients are related to primordial perturbation as follows:
\begin{eqnarray}
a_{lm}&=&4\pi (-\imath)^l \int \frac{d^3\mathbf k}{(2\pi)^3} \Phi(\mathbf k)\,g_{l}(k)\,Y^*_{lm}(\hat {\mathbf k}),\label{alm}
\end{eqnarray}
where $\Phi(\mathbf k)$ is primordial perturbation in Fourier space, and $g_{l}(k)$ is a radiation transfer function.
For a Gaussian model for primordial perturbation, decomposition coefficients satisfy the following statistical properties:
\begin{eqnarray}
\langle a_{lm} \rangle &=& 0, \\
\langle a^*_{lm} a_{l'm'} \rangle &=& C_l\,\delta_{ll'}\delta_{mm'},
\end{eqnarray}
where $\langle\ldots\rangle$ denotes the average over the ensemble of universes.
Given a standard cosmological model, Sach-Wolf plateau is expected at low multipoles \citep{Modern_Cosmology}:
$l(l+1) C_l\sim \mathrm{const}$.
\end{appendix}
\bibliographystyle{unsrt}
|
\section{Introduction}
Since the recognition that winds from the central stars of planetary nebulae (PNe)
play an important role in the shaping of these objects \citep[e.g.][]{K78,B87}, and that only
a small fraction of them show circular symmetry, the interest in PN morphology
has triggered an active field in both theoretical and observational
astronomy. Many studies have been devoted to classify
\citep[e.g.,][]{B87,SCM92,M96} and to model the basic morphologies observed
(circular, elliptical and bipolar) with noticeable success \citep[e.g.,][]{BPI87,H97}.
However, high resolution images have shown that the morphologies of PNe are
far from simple. Multiple shells, multipolar structures, highly collimated
outflows, microstructures and peculiar geometries are present in many PNe and
cannot be explained with simplistic models \citep[e.g.,][]{ST98,MAVG06,MPG09}.
Nowadays, it is accepted that highly collimated outflows play a crucial role
in the shaping of PNe \citep{ST98}. Nevertheless, other physical processes
are probably present, so that most likely the shaping of PNe is a result of many
processes acting at the same time \citep[e.g.,][]{BF02}. In order to
understand complex PNe, the first step is to identify the structural
components present and to define their nature. High-resolution, spatially resolved
spectroscopy combined with narrow-band imaging has demonstrated to be a
powerful tool to disentagle the structural components in PNe, to infer their
nature and to constraint the ejection processes involved in their
formation \citep{MS92,V99,V08,G08,V98,L00}.
Although at first glance NGC\,7354 looks like an elliptical PN, extense H$\alpha$,
[\ion{O}{3}]$\lambda5007$, [\ion{N}{2}]$\lambda6584$ and mid-infrared imaging and
spectroscopic data have revealed that it possess a more complex structure
\citep{S83,B87,H97,Ph09}. In particular, \citet{B87} described this nebula as
consisting of an inner halo, a thin bright rim, two spike-like tails, and
low-excitation patches projected onto the rim-halo interface. The large-scale
structures observed in NGC\,7354 are qualitatively well described using the
interacting winds theory \citep[e.g.][]{BPI87,M95,H07} but no deep analysis
has been carried out for the small-scale structures and their relationship with the
large-scale ones.
In this work we present a detailed observational study of the morphology,
internal kinematics, and physical and chemical properties of NGC\,7354. These
data allow us to discuss and model each of the components present in the
object and to suggest a possible scenario for the formation of this nebula.
\section{Observations and Results}
\subsection{Optical Imaging}
Narrow-band images of NGC\,7354 were obtained on 1997 July 24 with the Nordic Optical
Telescope (NOT)\footnote{The NOT is operated on the island of La Palma jointly by Denmark,
Finland, Iceland, Norway, and Sweden, in the Spanish Observatorio del Roque de los
Muchachos of the Instituto de Astrofisica de Canarias.} and the HiRAC camera equiped with
a Loral CDD of 2048$\times$2048 pixels and a plate scale of $0\rlap{.}{''}11$ pixel$^{-1}$.
Two narrow-band filters were used: [\ion{N}{2}]$\lambda6584$ ($\Delta \lambda=10$\,\AA) and
H$\alpha$ ($\Delta \lambda=10$\,\AA). The exposure time for both filters
was 900s. Seeing was about $0\rlap{.}{''}9$ during the observations.
Figure~\ref{fig1} shows a mosaic of our H$\alpha$ and [\ion{N}{2}] images,
including unsharp masking images in both filters constructed to show up both
the large- and small-scale structures in the nebula. In these images, we can
identify the main structures previously described by other authors: the outer
shell, the elliptical inner shell, the bright equatorial knots, and the two jet-like
features. In the following we will describe in more detail each of these structures as
well as new morphological details that have not been previously mentioned.
The outer shell looks like a round envelope in the high-contrast images but it
appears as a faint cylindrical structure in the low-contrast images. Its size is
$\simeq 33''\times29''$ with the major axis oriented at
position angle PA\,$\simeq15^\circ$. The outer shell is brighter in
H$\alpha$ than in [\ion{N}{2}] although in [\ion{N}{2}] several bright knots
are observed at the edges of it.
The inner shell present an elliptical shape with its major axis oriented at PA
$\simeq30^\circ$ and a major and minor axis length of $\simeq$ 21$'$ $\times$ 16$''$,
respectively. This structure is noticeable fainter in [\ion{N}{2}] than in
H$\alpha$. Moreover, the regions along the minor axis are particularly bright
in H$\alpha$. A closer inspection of the images shows that the polar regions
deviate from the elliptical shape and appear as two bubbles. This is particularly
noticeable in H$\alpha$. We will refer to these regions of the inner shell
as the polar caps .
The bright equatorial knots are observed in [\ion{N}{2}] but not in
H$\alpha$. They are mainly concentrated in two groups, East and West, and
along the equatorial plane of the outer and inner shells.
The two jet-like features are bright in [\ion{N}{2}] but much fainter in
H$\alpha$. The northern feature is oriented at PA$\simeq13^\circ$ and extends
$\simeq7''$. The southern feature is oriented at PA $\simeq205^\circ$, extends
$\simeq11''$, and appears narrower than the
northern one. We note that the orientation of these two features does not
coincide with each other nor with the orientation of the outer and inner shells.
In order to improve the view of NGC\,7354, we retrieved a Hubble Space Telescope (HST)
image from the MAST Archive\footnote{Some of the data presented in this paper were
obtained from the Multimission Archive at the Space Telescope
Science Institute (MAST). STScI is operated by the Association of Universities for Research
in Astronomy, Inc,, under NASA contract NAS5-26555. Support for MAST for non-HST data provided
by the NASA Office of Space Science via grant NAG5-7584 and by other grants
and contracts.} (Proposal ID:~7501; P.I.: A. Hajian; date of observation:~1998
July 21; filter F658N; exposure time 1000\,sec). Fig.~\ref{fig2} shows
this image. The morphology of the outer and inner shells is similar to that
observed in ground based images (Fig.\,1). The jet-like freatures appear as
cometary tails with a bright knot facing the central star and faint tails
directed outwards. They present a knotty structure, particularly the southern
one. The HST image resolve the bright equatorial knots into a series of small
knots and filaments embedded in diffuse emission.
\subsection{Radio continuum}
The $\lambda3.6$\,cm radio continuum observations were made with the Very Large Array (VLA)
of the NRAO\footnote{The National Radio Astronomy Observatory (NRAO) is operated by
Associated Universities Inc., under cooperative agreement with the National
Science Foundation.} on 1996 May 31 in the DnC configuration. The standard VLA continuum mode
with a bandwidth of 100\,MHz and two circular polarizations was employed. Phase center was
set at RA(2000.0)=22$^{\rm h}$40$^{\rm m}$20\fs1, DEC(2000.0)=+61\arcdeg17\arcmin06\farcs0.
Flux calibrator was 3C48 (adopted flux density 3.3 Jy) and phase calibrator
was 0019+734 (observed flux density 1.0 Jy). Total on-source integration time was 30 minutes.
The data were calibrated and processed using standard procedures of the
Astronomical Image Processing System (AIPS) package of the NRAO.
Fig.~\ref{fig3} shows an uniform-weighted map of NGC\,7354. The
emission presents a circular morphology and extends $\simeq$42\arcsec\,in
diameter. Two emission maxima are observed separated by $\simeq$10\arcsec\,and
oriented at PA$\simeq100\arcdeg$. These emission maxima coincide with the
H$\alpha$ bright regions observed along the minor axis of the inner shell.
From our data we derive a peak flux density of 72\,mJy\,beam$^{-1}$ at
position RA(2000.0)=22$^{\rm h}$40$^{\rm m}$20\fs44,
DEC(2000.0)=+61\arcdeg17\arcmin06\farcs8, and a total flux density
of $502\pm4$\,mJy. We note that our map is similar to that obtained
by \citet{T74}, being the flux density values obtained in both works consistent with
each other and with optically thin thermal emission.
Adopting a distance of 1.5\,kpc for the nebula \citep[][see Sec. 3.1]{S86} and following the
formulation by \citet{MH67}, we derive a mean electron density of $\simeq$ 710\,cm$^{-3}$ and an
ionized mass of $\simeq$ 0.22\,$M_\odot$.
\subsection{Low Resolution Spectroscopy}
Low resolution, long-slit spectra were obtained with the Boller \& Chivens spectrograph
mounted on the 2.1\,m telescope at the San Pedro M\'artir Observatory
(OAN-SPM)\footnote{The Observatorio Astron\'omico Nacional at San
Pedro M\'artir (OAN-SPM) is operated by the Instituto de Astronom\'{\i}a of the Universidad
Nacional Aut\'onoma de M\'exico.} during three observing runs: 2002 June 14, 2002 August 7,
and 2002 December 10 and 11. A CCD SITe with $1024\times1024$ pixels was used as a detector.
We have used a 400 lines/mm dispersion grating and a slit width of 2\arcsec\,giving a
spectral resolution (FWHM) of 7\AA. Spectra reduction was performed using
IRAF\footnote{The Image Reduction and Analysis
Facility (IRAF) is distributed by the National Optical Astronomy Observatories, which are
operated by the Association of Universities for Research in Astronomy, Inc., under
cooperative agreement with the National Science Foundation.}
standard procedures. We have used four slit positions to cover specific regions of the
nebula. Fig.~\ref{fig4} shows these slits, labelled A to D, and the regions extracted from
each long-slit spectrum overplotted on the unsharp masking [\ion{N}{2}] image.
Regions are denoted by a letter refering to the slit position followed by a sequence number
along the slit. In total, 21 regions in NGC\,7354 have been analyzed.
Table\,1 presents dereddened line fluxes, observed H$\beta$ flux and the
logarithmic extinction coeficiente $c{_{\rm H\beta}}$ for each region, the last has been
derived from the observed H$\alpha$/H$\beta$ ratio assuming case B recombination.
The fluxes have been dereddened aplying the extinction law by \citet{CCM89}. Electron
temperature ($T_e$) has been derived from the [\ion{O}{3}] and/or [\ion{N}{2}]
emission lines, while electron density ($N_e$) has been derived from the [\ion{S}{2}],
[\ion{Cl}{3}] and/or [\ion{Ar}{4}] emission lines, in both cases using the Five-Level Atom
Diagnostic Package NEBULAR \citep{dRDH87,SD94} from IRAF. Derived physical parameters and
their estimated errors are shown in Table~\ref{tbl-2}. Error
estimates take into account the readout and photon noise and they are
propagated along the calculation of physical quantities. Whenever possible,
$N_e$ values were calculated using the value of $T_e$ derived from
the corresponding high- or low-excitation ion, [\ion{O}{3}] or [\ion{N}{2}], respectively.
Extinction within the nebula, as indicated by $c_{\rm H\beta}$, shows a
slight increase from South to North with values ranging from $\simeq$\,1.7 to
$\simeq$\,2.4, respectively. Relatively high values of $c_{\rm H\beta}$ $\simeq$
2.4 are found in the low-excitation equatorial knots. These values can be
compared to those in the surroundings of the knots where the extinction
decreases up to $c_{H\beta}$ $\simeq$ 1.7. This result suggests that the
equatorial knots are denser and/or contain more dust than the rest of the nebula.
Electron density slightly increases inwards from the outer shell (regions B1 and C6) with
values of $\simeq$~1\,000\,cm$^{-3}$ to the inner regions of the inner shell (region
B4) where values of $\simeq$ 2\,400\,cm$^{-3}$ are found. The bright
equatorial knots are clearly denser that both the outer and inner shells
with an average density of $\simeq$\,2\,600\,cm$^{-3}$. For the jet-like
features, the electron density is relatively low with values around
$\simeq$\,1\,300\,cm$^{-3}$.
Electron temperature also seems to slightly increase from $\simeq$ 13,000\,K at the walls
of the outer shell (regions B1, B6 and B7) to $\simeq$ 15,000\,K at the center of
the inner shell (region B3). In the bright equatorial knots, electron
temperature ranges from $\simeq$~10,000 to $\simeq$\,12,400\,K. For the northern
jet-like feature, no [\ion{O}{3}] and [\ion{N}{2}] lines
were detected with a good signal-to-noise; in the southern jet-like feature,
we have obtained an electron temperature of $\simeq$ 12,000\,K (regions A5 and
D3).
Ionic and elemental abundances values are listed in Tables 3 and 4,
respectively. They have been obtained with the task IONIC in IRAF and using
the ionization correction factors by \citet{KB94}, respectively.
Table~\ref{tbl-4} also lists abundances in other objects for comparison purposes.
Small abundance variations, within the estimated uncertainties, are observed throughout
NGC\,7354. In general, we found that all our abundance determinations are consistent with
those of PNe, see Table~\ref{tbl-4}. Since neither He nor N overabundance is observed,
as in the case of Type\,I PNe, we can say that it behaves like a Type\,II PN.
Elemental abundances for NGC\,7354 have been reported by several authors \citep{H97,MV02,P04,
S06}. However, a comparison of our abundance estimations with those from the literature is not
straightforward since we have obtained abundance values for several slit positions and specific
regions along them. In the case of \citet{H97}, they report abundance values of six regions
along a slit position very similar to our slit A. Their abundances tend to be two or three times
higher than ours in the case of O/H, N/H and S/H but lower in He/H and similar in Ar/H.
Comparison with other studies can only be done through average values along each of our slit
positions or the total average obtained from all our studied regions. We have compared our total
averaged O/H, N/H, S/H, Ar/H abundances with those from \citet{P04} and \citet{S06}. In both
studies, their abundances seem to be lower than ours but in the case of \citet{S06} this difference
may be due to tha fact that they consider the PN as a whole and excluded special features in the
nebula.
\subsection{High Resolution Spectroscopy}
High-resolution, long-slit spectra were obtained in 2002 July 15 to 17 and 2007
July 10 to 17 with the Manchester Echelle Spectrometer (MES; \citep{M03}) mounted on the
2.1\,m telescope at San Pedro M\'artir Observatory (OAN-SPM). A Site CCD with
1024$\times$1024 pixels was used as a detector. Binnings of $1 \times
1$ and $2 \times 2$ were used in 2002 and 2007, respectively. Slit width was
1.6\arcsec\, and the achieved spectral resolution (FWHM) is 12 km
s$^{-1}$. The slit was oriented North-South and centered at several right
ascensions across the nebula, except in those slits that covered the jet-like
features, which have been centered on the central star and oriented at PAs
13$^{\circ}$ and 25$^{\circ}$. Figure\,5 shows the used slit positions,
numbered from 1 to 8, superimposed on the unsharp masking [\ion{N}{2}]
image. Data reduction was carried out with the IRAF package.
Grey-scale, position-velocity (PV) maps, derived from the eight long-slit
spectra, are shown in Fig.~\ref{fig6} for the H$\alpha$, [\ion{N}{2}]$\lambda6584$, and
\ion{He}{2}$\lambda6560$\ emission lines. From the long-slit spectra, we
derived a heliocentric systemic velocity of $-42\pm2$\,km\,s$^{-1}$ for
NGC\,7354, in excellent agreement with $-41\pm2$\,km\,s$^{-1}$ deduced by \citet{S83}.
Through this paper, we will consider the systemic velocity as the origin for quoting internal
radial velocities and the declination of the central star as the origin for quoting distance
measurements. The PV maps allow us to recognize the structural components that
have been identified in the direct images. In the following, we will describe
the spatio-kinematical properties of these components.
The [N\,{\sc ii}] emission from the outer shell is recognizable at slit positions
1,2,3,5, and 6. The emission is very faint and in the central nebular regions it appears
superposed by the stronger emission from the inner shell. Maximum radial
velocity of $\simeq$ 35 km\,s$^{-1}$ is observed at the center and decreases with
distance to the central star. Emission from the outer
shell can be recognized in the H$\alpha$ line (Fig.\,6) by its spatial
extend. However, the large thermal width and, probably, low expansion velocity
in this line do not allow us to obtain detailed kinematic information. The
outer shell cannot be recognized in the long-slit spectra of the
\ion{He}{2}$\lambda6560$ line.
The inner shell can be identified at all slit positions in the three
lines. The emission feature of the inner shell appears as a velocity ellipse
in the PV maps. The size of the velocity ellipse is maximum at slit 8 with
values of 30$''$ in H$\alpha$ and [\ion{N}{2}], and 20$''$ in
\ion{He}{2}. Maximum velocity splitting is observed at the center of the
nebula (slits 3, 7, and 8) and amounts 60, 56 and 48 km\,s$^{-1}$ in
H$\alpha$, [\ion{N}{2}], and \ion{He}{2}, respectively. This implies expansion
velocities for the inner shell of 30, 28 and 24 km\,s$^{-1}$,
respectively. Since we estimate an error of $\leq 1$\,km\,s$^{-1}$
in our velocity measurements, these results show that the \ion{He}{2} emission is
confined to a slow expanding, inner thin layer of the inner shell while the
[\ion{N}{2}] traces the outer layer that expands faster.
It is worth noticing the very faint and wide emission observed in [\ion{N}{2}]
(slit positions 2, 3, 7 and 8) located at about 12\arcsec\,and 8\arcsec\,to
the North and South of the central star, respectively. The velocity of this faint emission
spreads from $\simeq -45$ to $\simeq33$\,km\,s$^{-1}$. Comparing our [\ion{N}{2}]
direct image and these particular PV maps, we identify these weak emissions with the
two polar caps located on the main symmetry axis of the inner elliptical shell.
All slit positions in our [\ion{N}{2}] PV maps, show the presence of at least one of the
bright low-ionization knots identified in our optical [\ion{N}{2}] images (Fig.~\ref{fig1},
right panel). In all these PV maps we can see that most of them are located very close to the
zero position line, i.e. almost on the equatorial plane of the nebula. Although these PV maps
show that emission arising from different bright knots may be mixed due to projection
effects, we can estimate that their expansion velocity ranges between the inner and outer
shell expansion velocities, 28\,km\,s$^{-1}$ to $\simeq$\,40\,km\,s$^{-1}$, with only two of
them having lower velocities, $\simeq$\,24\,km\,s$^{-1}$. A final remark on the low-ionization
knots is that in our [\ion{N}{2}] PV maps, slit positions 1 and 6, we can see two weak spots
of emission almost simetrically located at about 14\arcsec\,to the North and South from the
central star, respectively. Both emission spots show negative and very low velocities of
$\simeq\,-5$\,km\,s$^{-1}$. We identify these two spots of emission as coming from two bright
knots that, in projection, appear to be located on the outer shell wall (see Fig.~\ref{fig1},
[\ion{N}{2}] images).
Although [\ion{N}{2}] emission from the two jet-like features can be identified at slit
positions 2 and 5 (Fig.~\ref{fig6}), it can be better analyzed at slit positions 7 and 8
which were specially selected to cover these structures. In the corresponding PV maps we
can see the conspicuous emission arising from the two jet-like features. While the North
jet-like feature emission is slightly redshifted, the South feature is slightly blueshifted.
Both features show a small radial velocity, between 0 and 5 km s$^{-1}$ which suggest that
they are moving almost in the plane of the sky.
\section{Discussion}
\subsection{Morphokinematic structure and modeling}
As we have described, NGC\,7354 shows four different structures: two large scale
structures, the outer and the inner shells, the last having two polar caps located on its
symmetry axis; a number of low-ionization bright knots lying about the equatorial plane, and
two jet-like features located close to the North and South of the nebula, slightly inclined
respect to the inner shell symmetry axis. Our high dispersion spectra
indicate that (a) the inner shell is expanding with a velocity which is smaller than that of
the outer shell, (b) most of the bright equatorial knots are moving at velocities closer to
the outer shell, and (c) the jets' projected velocity is very small.
In order to test the overall morphology and kinematic structure of the nebula we have used
the interactive three-dimensional (3D) modeling tool {\sc shape} Ver. 2.0 \citep{SL06}.
{\sc shape} produces synthetic images and PV diagrams which can be compared directly with our
observed CCD images and PV maps. We have considered two ellipsoidal structures (inner and
outer shells), two spherical sections (polar caps), several small spheres with different
diameter (bright knots) and two geometrically thin cylinders of 2$''$ in diameter (jet-like
features). All these geometrical structures were slightly modified (smoothered, lengthened and/or
rounded) in order to match the overall looking of our [\ion{N}{2}] images. Besides, all the
structures were placed on the proper position to get the corresponding observed radial
velocity from our PV maps. Fig.~\ref{fig7} shows all the geometrical components used to
construct our {\sc shape} model.
Each one of the main structures was analized separately in order to obtain its spatial
velocity independently. Due to projection effects we are unable to distinguish neither if the
low-ionization knots are located between the inner and outer shell nor if they are lying on
the foreground or on the background side of the nebula. Thus, we have assumed that
the bright knots are located on the outer shell wall, according to their individual velocity
shown in our PV map. It is important to remark that the inclination angle of the
different structures is unknown and {\sc shape} does not solve it. Final geometrical and
kinematic parameters are shown in Table~\ref{tbl-5}. Some of the inner shell basic parameters
can be compared to those found by \citet{H07} with an extended prolate ellipsoidal shell model.
While our derived inclination and position angles are consistent with the ones determined by
\citet{H07}, our expansion velocity value is larger than theirs. However, as they have
found in all the cases examined in their study, the [\ion{N}{2}] gas has a larger expansion
velocity than the [\ion{O}{3}] gas. Then, since we have determined the expansion velocity
through the [\ion{N}{2}] line emission, a difference in expansion velocities is expected.
Distance estimations found in the literature lie around 1.2\,kpc \citep{CKS92,Z95,Ph04} being
1.5\,kpc the largest estimation \citep{S86} and 0.88\,kpc the smallest one \citep{D82}.
We have adopted the largest value of 1.5\,kpc in order to derive an upper limit for the
kinematical ages. However, using a distance value of 1.2\,kpc implies a decrease in
the kinematical age estimation of only 20\%. But most important, distance uncertainty does
not modify the proposed chronological sequence of structure formation in our model.
From this model, we have obtained synthetic images and PV diagrams
which strongly resemble the observed ones, see Fig.~\ref{fig8} for an example of this good
match. Thus, from the proposed 3D model we have not only reproduced qualitatively the overall
morphology of NGC\,7354 but we have also quantitatively reproduced the kinematic behaviour of
each structure.
\subsection{Formation of NGC\,7354}
At present, binary star models have been succesful in explaining the origin and process
responsible for the equatorial density enhacement required in the interacting stellar wind
(ISW) model, or rather the generalized ISW (GISW) model, to explain early- and middle-type
elliptical and even bipolar large scale structures observed in PNe
\citep[e.g.][]{BPI87,M95,MF95}. On one hand, common envelope (CE) evolution in close binaries
has been proposed to be responsible for the expanding slow wind torus needed in
GISW \citep[e.g][]{RL96,F96,TT96,S98}. On the other hand, accretion disks in binary systems
seem to account for the narrow waist bipolar morphologies and even the jets present in PNe.
Details may vary on each binary model in order to explain individual morphologies and
structures, but in all of them the main idea is that of a jet launched by the central star,
or its companion \citep[][among others]{SR00,GAF04,S07,D08,AS08}.
Since at first glance NGC\,7354 looks like a rather elliptical nebula, one
may think that the single star models (ISW or GISW) can explain the presence of the outer
and inner shells, and even the formation of the two jet-like features
observed in our [\ion{N}{2}] image \citep{BPI87}. However, it is unable to explain why all
the different structures show different position angles (PA's) on the sky. This
characteristic may indicate that the direction of ejection varies with time, and more likely
that all the features were formed as independent events. Although nowadays there is no
evidence of NGC\,7354 possessing a binary nucleus, one may turn around and take a look at
binary star models to try to understand in a consistent way the observed morphology of
NGC\,7354.
It has been proposed that the GISW model coupled with the predictions of the CE evolution
theory can account for the formation of elliptical and even bipolar PNe morphologies
\citep{SL94}. Then, we could succesfully explain the outer and inner shells in NGC\,7354
based on this mixed model in the following way. After the CE phase has ended, i.e. the
envelope has been ejected, the binary may become closer, mass transfer take place, an
accretion disk is formed and a jet may be launched \citep{SL94}. If we assume that the
subsequent evolution is similar to those described in the models for pPNe \citep{SR00,LS03,
D08,AS08} it is possible that a jet arise from an accretion disk. Then, the accretion disk
theory would be able to explain also the two jet-like features observed in NGC\,7354.
Moreover, some of these models even propose the formation of an expanding dense ring in the
equatorial plane \citep{SR00} which in the case of NGC\,7354 might be related to the
equatorial bright knots. As the nebula evolve, a combination of the processes and mechanisms
just mentioned might take place.
With this in mind, we suggest a possible qualitative scenario for the formation of the main
structures present in NGC\,7354. A slow wind is lost by a central binary system ongoing a CE
phase, producing the mild elliptical outer shell. When the AGB fast wind begins, it
finds an equatorial density enhaced enviroment forming the bright elliptical rim or inner
shell. At this point the CE has been completely ejected and mass transfer from the secondary
(main-sequence star) into the primary (white dwarf) will form an accretion disk and
eventually a couple of jets will be launched. Once the jets have appeared, they will ``push''
the polar ends of the inner shell, detaching them from it. Eventually, the two jets would
break the caps and escape from the main body of the nebula. Along with the processes
mentioned, the binary system may be precessing and each of the structures would show a
different PA on the sky. In the case of the two jets, precession may cause that they impinge
on the polar caps with a certain angle, i.e. not along the symmetry axis of the caps,
breaking them apparently on the ``base'' of them. According to our derived kinematical ages,
the chronological sequence of formation seem to be consistent with the above description.
The outer shell is the oldest structure present in the nebula with an age of about 2500\,yr.
The next younger structure would be the inner elliptical shell with $\simeq$1600\,yr.
Kinematical ages for the two jet-like features indicate that they are coeval and both were
formed almost at the same time as the inner elliptical shell. However, we should notice that
the age was derived assuming that they have been moving with the same velocity (60 km\,s$^{-1}$)
since their launch. According with the theory of formation of jets this low initial velocity would
be very unlikely, since they arise from accelerated flowing material. Then, more likely the
jets possessed a higher initial velocity but they have been restrain along its way. Maybe
the main cause of this decceleration was the interaction of the jets with the polar caps.
Regarding the equatorial low-excitation bright knots, as we have mentioned above they may be
related to the dense expanding ring formed around the binary system. The bright knots show
velocities ranging between the inner and outer shell expansion velocities. This may indicate
that they are moving with a similar velocity as the particular ambient gas
in which they are embedded. We may compare these bright knots with the microstructures
observed in NGC\,2392 \citep[eskimo nebula,][]{ODell02} and NGC\,7662 \citep{PPB04}. Although
each one of these nebulae is seen with a different view angle (NGC\,2392 is seen pole-on and
NGC\,7662 is seen edge-on) both of them show various low-ionization knots located beyond the
inner structure and mainly on the outer envelope. If we compare the general morphology of
NGC\,7354 with that observed in NGC\,2392 and NGC\,7662 one can see that the three of them
show an ellipsoidal inner structure surrounded by on outer envelope with low-ionization
bright knots located on the outer structure plus jet-like features, interpreted as FLIERs in
the case of NGC\,7662. In the case of NGC\,7354 since it is observed with an edge-on view
angle and it is inclined with respect to the plane of the sky, we are not able to observe the
real morphology of the equatorial brights knots.
However, one can imagine that if one could get a pole-on view, i.e. with the inner
shell major axis aligned with our line of sight, one would see bright knots located in a zone
between the inner and outer shell very alike the arragement of knots with tails observed in
the eskimo nebula. Therefore, having all
these elements together, we suggest that, as in the cases of NGC\,7662 and NGC\,2392, the
equatorial bright knots observed in NGC\,7354 are not spherical structures but knots with
tails, or even filamentary structures. This last suggestion is strongly supported by the high
resolution [\ion{N}{2}] HST image, Fig.~\ref{fig2}.
A full model with MHD simulations is beyond the scope of the present study, but we are aware
that it could clarify some aspects of our work.
\section{Conclusions}
We have carried out a detailed morphological and kinematical analysis of the planetary nebula
NGC\,7354. In addition, we have derived the physical conditions and elemental abundances in 21
regions of the nebula. The main conclusionas of this work can be sumarized as follows.
Physical parameters of all the different structures show that there are slight variations, both
in electron density and electron temperature, within the nebula. Considering our error estimates,
while density seems to slightly increase inward from the outer shell border to the interior of
the inner elliptical shell, temperature values seem to slightly increase in the same direction.
Equatorial bright knots are clearly denser that both shells with temperatures in the typical
range for ionized gas. Both jet-like features present a quite low density value and
temperature was only determined for the South jet-like feature. All our derived physical
parameter values are consistent with the ones expected for radiatively excited gas. Local
extinction values within the nebula show a slight increasing gradient going from South to
North of the nebula and from West to East along the location of the equatorial knots.
Based on our kinematical data we have obtained a model that consistently reproduce the
overall as well as the detailed morphology and PV maps of the nebula, using the
3D interactive tool {\sc shape}. The final geometrical components included in the model are:
i) two shells with similar inclination angle (respect to our
line of sight) but different position angle of the projected semi-major axis;
ii) two semispherical caps placed on the top of the inner elliptical shell;
iii) thirteen spherical knots with different sizes, whose velocities spanned between the
corresponding velocity values for the outer and inner shells, placed at different positions
on the outer shell surface according to the optical image and observed spectra; and
iv) two geometrically thin cylinders corresponding to North and South jet-like features.
Finally, although there is no evidence of binarity in NGC\,7354 at present, we suggest a
qualitative scenario for the formation of the different structures in the nebula
based on the theory of CE evolution and formation of accretion disks in binary
systems found in the literature.
\acknowledgments
This project has been supported by grants from CONACYT (49002, 45848) and PAPIIT-UNAM
(IN111903, IN109509). LFM acknowledge support from grants AYA2005-01495 of the Spanish MEC
(co-funded by FEDER funds), AYA2008-01934 of the Spanish MICINN (co-funded by FEDER funds),
and FQM1747 of the Junta de Andaluc\'{\i}a. LO aknowledge CONACYT for his posdoctoral research
scholarship. SZ acknowledge support from the UNAM-ITE colaboration agreement 1500-479-3-V-04.
We are grateful to the staff of all the astronomical facilities and systems used in this
research, in particular to Mr. Gustavo Melgoza for assistance during OAN-SPM observations.
|
\section{ Introduction }
One of the most extensively used ways to describe fundamental particles and
fields is still a theory of relativistic wave equations (RWE), the
foundations of which have been laid by Dirac [1], Fierz and Pauli [2,3],
Bhabha [4,5], Harish-Chandra [6,7], Gel'fand and Yaglom [8], Fedorov [9,10].
This theory has been advanced proceeding from the assumption that a
relativistic-invariant description of both massive and massless particles
(fields) may always be reduced to a system of the first-order differential
equations with constant factors, in the matrix form being given as follows:
\begin{equation}
\label{eq1.1}
\left( {\gamma _\mu \partial _\mu + \gamma _0 } \right)\psi
\left( x \right) = 0 \quad \left( {\mu =1\div 4} \right).
\end{equation}
Here $\psi \left( x \right)$ is multicomponent wave function transformed in
terms of some reducible Lorentz group representation $T$, $\gamma _\mu $ and
$\gamma _0 $ are square matrices.
In the case when the matrix $\gamma _0 $ is nonsingular $\left( {\det \gamma
_0 \ne 0} \right)$, equation (\ref{eq1.1}) describing a massive particle may be
reduced to the following form by multiplication into $m \gamma _0 ^{ - 1}$ :
\begin{equation}
\label{eq1.2}
\left( {\gamma _\mu \partial _\mu + m I} \right) \psi
\left( x \right) = 0 ,
\end{equation}
where $m$ is a parameter associated with mass, $I$ is unity matrix.
A choice of the matrices $\gamma _\mu $ in equations (\ref{eq1.1}) and
(\ref{eq1.2})
is limited
by the following requirements (e.g., see [8,9]):
\noindent
~~i) invariance of the equation with respect to the transformations of its own
Lorentz group;
\noindent
~ii) invariance with respect to reflections;
\noindent
iii) possibility for derivation of the equation from the variational principle.
Equations of the form (\ref{eq1.2}) meeting requirements i)--iii) are known as
relativistic wave equations (RWE); equations of the form (\ref{eq1.1}) with the same
requirements are known as generalized RWE [9].
From requirement i) and from the condition of theory's irreducibility with
respect to the Lorentz group it follows that the function $\psi $ must be
transformed by some set of linking irreducible Lorentz-group
representations, forming what is known as a scheme for linking. The
representations $\tau \sim \left( {l_1 , l_2 } \right)$ and
${\tau }' \sim \left( {{l}'_1 ,{l}'_2 } \right)$ are referred to
as linking if ${l}'_1 = l_1 \pm \frac{1}{2}, \quad {l}'_2
= l_2 \pm \frac{1}{2}.$
Aside from a choice of the wave function $\psi $, in definition of different
spin and mass states possible for the particle described by equations (\ref{eq1.1}) and
(\ref{eq1.2}) the matrices $\gamma _4 $ and $\gamma _0 $ are of particular
importance. Properties of the matrix $\gamma _4 $ are discussed
comprehensively in [8]. A structure of the matrix $\gamma _0 $ is determined
in [5,9]. Specifically, requirement i) results in reducibility of $\gamma _0
$ to the diagonal form, the matrix being composed of independent scalar
blocks corresponding to the irreducible representations of $\tau $. For
$\det \gamma _0 = 0$ some of these blocks are zero. As follows from
requirement ii), nonzero elements $a_\tau $ of the matrix $\gamma _0 $
satisfy the relation
\begin{equation}
\label{eq1.3}
a_\tau = a_{\dot {\tau }} ,
\end{equation}
where $\dot {\tau }$ is representation conjugate to $\tau $ with respect to
the spatial reflection, i.e., if $\tau \sim \quad \left( {l_1 ,l_2 }
\right)$, we have $\dot {\tau } \sim \left( {l_2 ,l_1 } \right)$.
In case of the finite-dimensional representations requirement iii) also leads
to the relation of (\ref{eq1.3}).
A distinctive feature of most well-known RWE of the form (\ref{eq1.2}) (Dirac
equation for spin $\frac{1}{2}$, Duffin-Kemmer equations for spins 0 and 1,
Fierz-Pauli equation for spin $\frac{3}{2})$ is the fact that they involve a
set of the Lorentz group representations \textit{minimally necessary}
for framing of a theory of this spin.
Such an approach in the case of$\det \gamma _0 = 0$ results in equations for
zero-mass particles (e.g., Maxwell equations). Because of this, selection of
$\det \gamma _0 = 0$ (also including $\gamma _0 = 0)$ in a theory of RWE is
associated with a description of massless particles [9,11].
It is known that, as distinct from the description of massive particles, in
a theory of massless particle with integer spin some of the wave-function
components are unobservable (potentials) and others - observable
(intensities). In consequence, for the potentials one can define the gauge
transformations and impose additional requirements excluding ``superfluous''
components of $\psi $. But for the description of massive particles by RWE
reducible to the form given by (\ref{eq1.2}),
the above-mentioned differentiation of
the wave-function components is not the case. In other words, the notion of
the gauge invariance of RWE (\ref{eq1.1}) in the sense indicated previously is
usually used for massless theories.
At the same time, there are papers, where so-called
\textit{massive gauge-invariant theories} are considered taking
other approaches. Illustrative examples are furnished by St\"uckelberg's
approach to the description of a massive spin 1 particle (see [12] and
references herein) and by a
$\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\frown$}}\over {B}}
\wedge \mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\frown$}}\over
{F}} $-theory [13--16] claiming for the description of string interactions
in 4-dimensional space and suggesting a mechanism
(differing from Higgs's) of the mass
generation due to gauge-invariant mixing of electromagnetic and massless
vector fields with zero helicity. In the literature this field is called the
Kalb-Ramond field [15,16] and the notoph [17]. Because of this, one should
clear the question concerning the status of massive gauge-invariant fields
in the theory of RWE.
Another feature of well-known RWE is the fact that on going from equation of
the form (\ref{eq1.2}) for a massive spin $S$ particle
to its massless analog of the
form (\ref{eq1.1}), by making the substitution $mI \to \gamma _0 ,\det \gamma _0
= 0$, not all of the helicity values from $ + S$ to $ - S$ are retained, a
part of them is lost. This is the case when passing from the Duffin-Kemmer
equation for spin 1 to Maxwell equations with the dropped-out zero helicity.
In some modern models there is a necessity for simultaneous description of
different massless fields [18]. Within the scope of a theory of RWE, it
seems possible to solve this problem by the development of a scheme for
passage from (\ref{eq1.2}) to (\ref{eq1.1})
RWE with the singular matrix $\gamma _0$
retaining not only maximal but also intermediate helicity values.
By authors' opinion, solution of the stated problems is important
considering a possibility of using the well-developed apparatus of a theory
based on RWE in modern theoretical field models including the
phenomenological description of strings and superstrings in a space of the
dimension $d = 4$.
\section{GAUGE-INVARIANT THEORIES FOR MASSIVE SPIN 0 AND 1 PARTICLES}
Let us consider the following set of the Lorentz group irreducible
representations in a space of the wave function$\psi $
\begin{equation}
\label{eq2.1}
\left( {0,0} \right)\oplus \left( {\frac{1}{2},\frac{1}{2}}
\right) \oplus \left( {0,1} \right) \oplus \left( {1,0}
\right).
\end{equation}
The most general form of the corresponding (\ref{eq2.1}) tensor system of the
first-order equations meeting the requirements i) -- iii) is given by
\begin{subequations}
\label{eq2.2}
\begin{eqnarray}
\label{eq5}
\alpha \partial _\mu \psi _\mu + a\psi _0 =0,
\\
\label{eq6}
\beta ^ * \partial _\nu \psi _{\mu \nu } + \alpha^*
\partial _\mu \psi _0 + b\psi _\mu = 0,
\\
\label{eq7}
\beta \left( { - \partial _\mu \psi _\nu + \partial _\nu \psi
_\mu } \right) + c\psi _{\mu\nu } = 0.
\end{eqnarray}
\end{subequations}
Here $\psi _0 $ is scalar, $\psi _\mu $ is vector, $\psi _{\mu \nu } $ is
antisymmetric second-rank tensor; $\alpha , \beta $ are arbitrary
dimensionless, generally speaking, complex parameters, and
$a,b,c$ are real nonnegative parameters, the dimension of which
on selection of $\hbar = c = 1$ is coincident with that of mass
(massive parameters). Writing system (\ref{eq2.2}) in the matrix form (\ref{eq1.1}), we
obtain in the basis
\begin{equation}
\label{eq2.3}
\psi = \left( {\psi _0,\psi _\mu ,\psi _{\mu \nu
} } \right) - column
\end{equation}
for the matrix $\gamma _0 $ the following expression:
\begin{equation}
\label{eq2.4}
\gamma _0 = \left( {{\begin{array}{*{20}c}
a \hfill & \hfill & \hfill \\
\hfill & {b I_4 } \hfill & \hfill \\
\hfill & \hfill & {c I_6 } \hfill \\
\end{array} }} \right).
\end{equation}
(Matrices of the form $\gamma _\mu $ are not given as they are of no use for
us in further consideration.)
In the general case, when none of the parameters in (\ref{eq2.2}) is zero, this
system describes a particle with a set of spins 0, 1 and with two masses
\begin{equation}
\label{eq2.5}
m_1 = \frac{\sqrt {a b} }{\left| \alpha
\right|},\quad m_2 = \frac{\sqrt {b c} }{\left| \beta \right|},
\end{equation}
the mass $m_1 $ being associated with spin 0 and $m_2 $ with spin 1.
Omitting cumbersome calculations, we will verify this during analysis of
special cases.
Imposing on the parameters of system (\ref{eq2.2}) the requirement
\begin{equation}
\label{eq2.6}
\frac{\sqrt a }{\left| \alpha \right|} = \frac{\sqrt c }{\left|
\beta \right|},
\end{equation}
we obtain RWE for a particle with spins 0, 1 and one mass $m = m_1 = m_2 $.
At $\alpha = 0$ system (\ref{eq2.2}) goes to the Duffin-Kemmer equation of
a particle with spin 1 and mass $m = m_2 $
\begin{subequations}
\label{eq2.7}
\begin{eqnarray}
\label{eq12}
\beta ^ * \partial _\nu \psi _{\mu \nu } + b\psi _\mu
= 0,
\\
\label{eq13_1}
\beta \left( { - \partial _\mu \psi _\nu + \partial _\nu \psi
_\mu } \right) + c\psi _{\mu \nu } = 0.
\end{eqnarray}
\end{subequations}
Finally, by setting in (\ref{eq2.2}) $\beta = 0$, we arrive at the
Duffin-Kemmer equation for a particle with spin 0 and mass $m=m_1$:
\begin{subequations}
\label{eq2.8}
\begin{eqnarray}
\label{eq13}
\alpha \partial _\mu \psi _\mu + a\psi _0 =0,
\\
\label{eq14}
\alpha^* \partial _\mu \psi _0 + b\psi _\mu =0.
\end{eqnarray}
\end{subequations}
Now we consider the case that is of great interest for us, when the
parameters $a,b,c$ determining a structure of the matrix $\gamma _0 $
in (\ref{eq2.4}) are selectively set to zero.
In system (\ref{eq2.2}) setting
\begin{equation}
\label{eq2.9}
a = 0,
\end{equation}
we have the following system of equations:
\begin{subequations}
\label{eq2.10}
\begin{eqnarray}
\label{eq17}
\partial _\mu \psi _\mu = 0,
\\
\label{eq18}
\beta ^ * \partial _\nu \psi _{\mu \nu } + \alpha ^ *
\partial _\mu \psi _0 + b\psi _\mu = 0,
\\
\label{eq19}
\beta \left( { - \partial _\mu\psi _\nu + \partial _\nu \psi
_\mu } \right) + c\psi _{\mu \nu } = 0,
\end{eqnarray}
\end{subequations}
that, being written in the matrix form of (\ref{eq1.1}),
corresponds in basis (\ref{eq2.3})
to the singular matrix $\gamma _0 $
\begin{equation}
\label{eq2.11}
\gamma _0 = \left( {{\begin{array}{*{20}c}
0 \hfill & \hfill & \hfill \\
\hfill & {b I_4 } \hfill & \hfill \\
\hfill & \hfill & {c I_6 } \hfill \\
\end{array} }} \right) .
\end{equation}
From system (\ref{eq2.10}) one can easily derive the second-order equations
\begin{eqnarray}
\label{eq2.12}
\Box\psi _0 = 0
\\
\label{eq2.13}
\Box\psi _\mu - \frac{c\alpha ^ * }{\left| \beta
\right|^2}\partial _{\mu } \psi _0 - \frac{b c}{\left| \beta
\right|^2}\psi _\mu = 0.
\end{eqnarray}
As regards the scalar function $\psi _0 $ governed by equation (\ref{eq2.12}), the
following aspects must be taken into account. System (\ref{eq2.10}) is invariant
with respect to the gauge transformations
\begin{equation}
\label{eq2.14}
\psi _0 \to \psi _0 - \frac{1}{\alpha ^ * }\Lambda
,\quad \psi _\mu \to \psi _\mu +
\frac{1}{b}\partial _\mu \Lambda ,
\end{equation}
where the gauge function $\Lambda $ is limited by the constraint
\begin{equation}
\label{eq2.15}
\Box\Lambda=0.
\end{equation}
From comparison between (\ref{eq2.15}) and (\ref{eq2.12}) it follows that the function $\psi
_0 $ acts as a gauge function and hence provides no description for a
physical field. In other words, gauge transformations (\ref{eq2.14}) and
(\ref{eq2.15}) make it
possible to impose an additional condition
\begin{equation}
\label{eq2.16}
\psi _0 = 0.
\end{equation}
In this case system (\ref{eq2.10}) is transformed to system (\ref{eq2.7}) describing a
massive spin 1 particle, whereas equation (\ref{eq2.13}), considered simultaneously
with (\ref{eq17}), goes to an ordinary Proca equation. In this way the gauge
invariance of system (\ref{eq2.10}), as compared to (\ref{eq2.2}), leads to a decrease in
physical degrees of freedom from four to three, exclusive of the spin 0
state.
Note that a similar result may be obtained without the explicit use of the
considerations associated with the gauge invariance. By the introduction of
\begin{equation}
\label{eq2.17}
\varphi _\mu = \psi _\mu + \frac{\alpha ^ *
}{b}\partial _\mu \psi _0
\end{equation}
system (\ref{eq2.10}) may be directly reduced to the form
\begin{subequations}
\label{eq2.18}
\begin{eqnarray}
\label{eq24}
\beta ^ * \partial _\nu \psi _{\mu \nu } + b\varphi _\mu = 0,
\\
\label{eq25}
\beta \left( { - \partial _\mu\varphi _\nu + \partial _\nu
\varphi _\mu } \right) + c\psi _{\mu \nu } =0
\end{eqnarray}
\end{subequations}
coincident with (\ref{eq2.7}).
This variant of a gauge-invariant theory is known [12] as a Stueckelberg
approach to the description of a massive spin 1 particle. We have considered
this variant for a complete study of the possibilities given by system
(\ref{eq2.2}).
In (\ref{eq2.2}) we set
\begin{equation}
\label{eq2.19}
c= 0.
\end{equation}
Then the initial system of equations (\ref{eq2.2}) takes the form
\begin{subequations}
\label{eq2.20}
\begin{eqnarray}
\label{eq27}
\alpha \partial _\mu \psi _\mu + a\psi _0 =0,
\\
\label{eq28}
\beta ^ * \partial _\nu \psi _{\mu \nu } + \alpha ^ *
\partial _\mu\psi _0 + b\psi _\mu = 0,
\\
\label{eq29}
- \partial _\mu \psi _\nu + \partial _\nu \psi _\mu =0.
\end{eqnarray}
\end{subequations}
According to (\ref{eq2.5}), it should describe a particle with the mass $m_1=
\frac{\sqrt {ab} }{\left| \alpha \right|}$ and with spin 0. By
convolution of equation (\ref{eq28}) with the operator $\partial _\mu $ we have
\begin{equation}
\label{eq2.21}
\Box\psi _0 + \frac{b}{\alpha ^ * }\partial _\mu \psi _\mu = 0.
\end{equation}
Comparing (\ref{eq2.21}) with (\ref{eq27}), we arrive at the equation
\begin{equation}
\label{eq2.22}
\psi _0 - \frac{ab}{\left| \alpha \right|^2}\psi _0 =0,
\end{equation}
that provides support for all the afore-said.
The states associated with spin 1, for the condition set by (\ref{eq2.19}),
disappear
due to the invariance of system (\ref{eq2.20}) with respect to the gauge
transformations
\begin{equation}
\label{eq2.23}
\psi _{\mu \nu } \to \psi _{\mu \nu } -
\frac{1}{\beta ^ * }\Lambda _{\mu \nu },\quad\psi _\mu
\to \psi _\mu + \frac{1}{b}\partial _\nu \Lambda
_{\mu \nu },
\end{equation}
where an arbitrary choice of the gauge function $\Lambda _{\mu \nu } $
is constrained by
\begin{equation}
\label{eq2.24}
\partial _\alpha \partial _\nu \Lambda _{\mu \nu } - \partial
_\mu \partial _\nu \Lambda _{\alpha \nu } = 0.
\end{equation}
On the other hand, as follows from equations (\ref{eq28}), (\ref{eq29}),
a similar equation
\begin{equation}
\label{eq2.25}
\partial _\alpha \partial _\nu \psi _{\mu \nu } - \partial
_\mu \partial _\nu \psi _{\alpha \nu } = 0
\end{equation}
is satisfied by the tensor $\psi _{\mu \nu } $\textbf{.} Consequently, a
choice of $\Lambda _{\mu \nu } $ is arbitrary enough to impose an
additional constraint
\begin{equation}
\label{eq2.26}
\partial _\nu \psi _{\mu \nu } = 0
\end{equation}
that is in accord with (\ref{eq2.25}). In this case system (\ref{eq2.20})
takes the form of
(\ref{eq2.8}), i.e. it actually describes a massive spin 0 particle.
Note also that system (\ref{eq2.20}) may be reduced to the form
\begin{subequations}
\label{eq2.27}
\begin{eqnarray}
\label{eq34}
\alpha \partial _\mu \varphi _\mu + a\psi _0 =0,
\\
\label{eq35}
\alpha ^{_ * }\partial _\mu \psi _0 + b\varphi _\mu =0
\end{eqnarray}
\end{subequations}
similar to (\ref{eq2.8}) by introduction of the vector
\begin{equation}
\label{eq2.28}
\varphi _\mu = \psi _\mu + \frac{\beta ^ *
}{b}\partial _\nu \psi _{\mu \nu }.
\end{equation}
Thus, the considered variant of a massive gauge-invariant theory is some
kind of an analog for the St\"uckelberg approach but applicable to the
description of a spin 0 particle. The authors have not found any mentioning
of such a description in the literature available.
In the formalism of RWE (\ref{eq1.1}) this theory is consistent with the matrix
$\gamma _0 $ of the form
\begin{equation}
\label{eq2.29}
\gamma _0 = \left( {{\begin{array}{*{20}c}
a \hfill & \hfill & \hfill \\
\hfill & {b I_4 } \hfill & \hfill \\
\hfill & \hfill & {0_6 } \hfill \\
\end{array} }} \right).
\end{equation}
Next we consider a set of the Lorentz group representations
\begin{equation}
\label{eq2.30}
\left( {\frac{1}{2}, \frac{1}{2}} \right) \oplus \left(
{\frac{1}{2}, \frac{1}{2}} \right)^\prime \oplus \left( {0, 1}
\right) \oplus \left( {1, 0} \right) ,
\end{equation}
where the representation $\left( {\frac{1}{2} , \frac{1}{2}}
\right)^\prime $conforms to the pseudovector or to the absolutely
antisymmetric third-rank tensor. The most general form of a tensor system of
the first-order equations based on representation (\ref{eq2.30}) and meeting the
above-mentioned requirements i)--iii) is given by
\begin{subequations}
\label{eq2.31}
\begin{eqnarray}
\label{eq39}
\alpha \partial _\nu \psi _{\mu \nu } + a \psi _\mu
= 0 ,
\\
\label{eq40}
\beta \partial _\nu \tilde {\psi }_{\mu \nu } + b \tilde
{\psi }_\mu = 0 ,
\\
\label{eq41}
\alpha ^ * \left( { - \partial _\mu \psi _\nu + \partial _\nu
\psi _\mu } \right) + \beta ^ * \varepsilon _{\mu \nu \alpha
\beta } \partial _\alpha \tilde {\psi }_\beta + c \psi _{\mu
\nu } = 0 .
\end{eqnarray}
\end{subequations}
Here $\tilde {\psi }_{\mu \nu } = \frac{1}{2} \varepsilon _{\mu
\nu \alpha \beta } \psi _{\alpha \beta } , \quad \tilde {\psi }_{\mu
} = \frac{1}{6} \varepsilon _{\mu \nu \alpha \beta }
\psi _{\nu \alpha \beta } , \quad \varepsilon _{\mu \nu \alpha
\beta } $ is the Levi-Civita tensor ($\varepsilon _{1 2 3 4} = -
i)$, $\psi _{\nu \alpha \beta } $ is antisymmetric third-rank tensor,
$\alpha , \beta $ are still arbitrary dimensionless, generally speaking,
complex parameters, $a,b,c$ are mass parameters.
Writing system (\ref{eq2.31}) in the form (\ref{eq1.1}), where $\Psi = \left( {\psi _\mu
, \tilde {\psi }_\mu , \psi _{\mu \nu } } \right)$ is column, for the
matrix $\gamma _0 $ we get the expression
\begin{equation}
\label{eq2.32}
\gamma _0 = \left( {{\begin{array}{*{20}c}
{a I_4 } \hfill & \hfill & \hfill \\
\hfill & {b I_4 } \hfill & \hfill \\
\hfill & \hfill & {c I_6 } \hfill \\
\end{array} }} \right).
\end{equation}
Now we elaborate on massive gauge-invariant theories obtainable from (\ref{eq2.31})
by manipulations with the parameters $a, b, c$.
Let us take the case
\begin{equation}
\label{eq2.33}
a = 0 .
\end{equation}
In this case we have a system of equations
\begin{subequations}
\label{eq2.34}
\begin{eqnarray}
\label{eq44}
\partial _\nu \psi _{\mu \nu } = 0 ,
\\
\label{eq45}
\beta \partial _\nu \tilde {\psi }_{\mu \nu } + b \tilde
{\psi }_\mu = 0 ,
\\
\label{eq46}
\alpha ^ * \left( { - \partial _\mu \psi _\nu + \partial _\nu
\psi _\mu } \right) + \beta ^ * \varepsilon _{\mu \nu \alpha
\beta } \partial _\alpha \tilde {\psi }_\beta + c \psi _{\mu
\nu } = 0
\end{eqnarray}
\end{subequations}
that, when formulated as (\ref{eq1.1}), is associated with the singular
matrix $\gamma _0 $
\begin{equation}
\label{eq2.35}
\gamma _0 = \left( {{\begin{array}{*{20}c}
{0_4 } \hfill & \hfill & \hfill \\
\hfill & {b I_4 } \hfill & \hfill \\
\hfill & \hfill & {c I_6 } \hfill \\
\end{array} }} \right) .
\end{equation}
From (\ref{eq2.34}) we can obtain the second-order equations
\begin{eqnarray}
\label{eq2.36}
\Big( \Box - \frac{b c}{|\beta|^2}\Big)\tilde {\psi }_\mu =0,
\\
\label{eq2.37}
\partial _\mu \tilde {\psi }_\mu = 0 ,
\\
\label{eq2.38}
\Box\psi _\mu - \partial _\mu \partial _\nu \psi _\nu =
0 .
\end{eqnarray}
Equations (\ref{eq2.36}) and (\ref{eq2.37}) denote that system (\ref{eq2.34}) involves the description
of a massive spin 1 particle. As shown by equation (\ref{eq2.38}), system (\ref{eq2.34})
describes also a massless field with the potential $\psi _\mu $. The latter
allows for involvement of the gauge transformation
\begin{equation}
\label{eq2.39}
\psi _\mu \to \psi _\mu + \partial _\mu \Lambda
\end{equation}
($\Lambda $ is arbitrary function), with respect to which system (\ref{eq2.34}) and
equation (\ref{eq2.38}) are invariant. The indicated invariance means that this
massless field is a Maxwell-type field with helicity $\pm 1$.
In this manner the gauge-invariant system (\ref{eq2.34}) irreducible with respect to
the Lorentz group offers a simultaneous description of a massive spin 1
particle and of a massless field with helicity $\pm 1$. In other words, here
we deal with a massive-massless gauge-invariant theory rather than massive
theory, as is the case for (\ref{eq2.10}) and (\ref{eq2.20}).
A similar result may be obtained if we set in (\ref{eq2.34})
\begin{equation}
\label{eq2.40}
b = 0 .
\end{equation}
Then we have
\begin{equation}
\label{eqeq2.41}
\gamma _0 = \left( {{\begin{array}{*{20}c}
{a I_4 } \hfill & \hfill & \hfill \\
\hfill & {0_4 } \hfill & \hfill \\
\hfill & \hfill & {c I_6 } \hfill \\
\end{array} }} \right) ,
\end{equation}
and the second-order equations following from the corresponding first-order
system
\begin{subequations}
\label{eq2.42}
\begin{eqnarray}
\label{eq52}
\alpha \partial _\nu \psi _{\mu \nu } + a \psi _\mu
= 0 ,
\\
\label{eq53}
\partial _\nu \tilde {\psi }_{\mu \nu } = 0 ,
\\
\label{eq54}
\alpha ^ * \left( { - \partial _\mu \psi _\nu + \partial _\nu
\psi _\mu } \right) + \beta ^ * \varepsilon _{\mu \nu \alpha
\beta } \partial _\alpha \tilde {\psi }_\beta + c \psi _{\mu
\nu } = 0
\end{eqnarray}
\end{subequations}
are of the form
\begin{eqnarray}
\label{eq2.43}
\Big(\Box - \frac{a c}{\left| \alpha \right|^2}\Big)\psi _\mu =0,
\\
\label{eq2.44}
\partial _\mu \psi _\mu = 0 ,
\\
\label{eq2.45}
\Box\tilde {\psi }_\mu - \partial _\mu \partial _\nu \tilde {\psi
}_\nu = 0 .
\end{eqnarray}
Equation (\ref{eq2.45}) and system (\ref{eq2.42}) are invariant with respect to the gauge
transformations
\begin{equation}
\label{eq2.46}
\tilde {\psi }_\mu \to \tilde {\psi }_\mu + \partial
_\mu \tilde {\Lambda } .
\end{equation}
Thus, here we deal again with a gauge-invariant massive-massless spin 1
theory.
Let us consider another set of representations
\begin{equation}
\label{eq2.47}
\left( {0, 0} \right)^\prime \oplus \left(
{\frac{1}{2}, \frac{1}{2}} \right)^\prime \oplus \left( {0, 1}
\right) \oplus \left( {1, 0} \right) \quad ,
\end{equation}
where $\left( {0, 0} \right)^\prime $ is associated with the absolutely
antisymmetric fourth-rank tensor $\psi _{\mu \nu \alpha \beta } $. The
most general tensor formulation of RWE based on the set of representations
given in (\ref{eq2.47}) takes the form
\begin{subequations}
\label{eq2.48}
\begin{eqnarray}
\label{eq58}
\alpha \partial _{\left[ \mu \right.} \psi _{\nu \alpha \beta \left.
\right]} + a \psi _{\mu \nu \alpha \beta } =
0 ,
\\
\label{eq59}
\alpha ^ * \partial _{\left[ \nu \right.} \psi _{\alpha \beta \left.
\right]} + \beta ^ * \partial _\mu \psi _{\mu \nu
\alpha \beta } + b \psi _{\nu \alpha \beta } =
0 ,
\\
\label{eq60}
\beta \partial _\nu \psi _{\nu \alpha \beta } + c \psi
_{\alpha \beta } = 0 ,
\end{eqnarray}
\end{subequations}
where the following notation is used:
\begin{eqnarray}
\label{eq2.49}
\partial _{\left[ \nu \right.} \psi _{\alpha \beta \left. \right]}
\equiv \partial _\nu \psi _{\alpha \beta } +
\partial _\beta \psi _{\nu \alpha } + \partial _\alpha
\psi _{\beta \nu } ,
\\
\label{eq62}
\partial _{\left[ \mu \right. } \psi _{\nu \alpha \beta \left. \right]}
\equiv \partial _\mu \psi _{\nu \alpha \beta } -
\partial _\nu \psi _{\mu \alpha \beta } + \partial
_\alpha \psi _{\mu \nu \beta } - \partial _\beta \psi _{\mu
\nu \alpha } .
\end{eqnarray}
After introduction into system (\ref{eq2.48}) of the dual conjugates $\tilde
{\psi }_{\mu \nu } , \tilde {\psi }_\mu $ and pseudoscalar $\tilde
{\psi }_0 = \frac{1}{4\mbox{!}} \varepsilon _{\mu \nu \alpha
\beta } \psi _{\mu \nu \alpha \beta } $ instead of the tensors $\psi
_{\mu \nu } , \psi _{\nu \alpha \beta } , \psi _{\mu
\nu \alpha \beta } $, it is conveniently rewritten to give
\begin{subequations}
\label{eq2.50}
\begin{eqnarray}
\label{eq63}
\alpha \partial _\mu \tilde {\psi }_{\mu } + a \tilde {\psi
}_0 = 0 ,
\\
\label{eq64}
\beta ^ * \partial _\nu \tilde {\psi }_{\mu \nu } +
\alpha ^ * \partial _\mu \tilde {\psi }_0 + b \tilde {\psi
}_\mu = 0 ,
\\
\label{eq65}
\beta \left( { - \partial _\mu \tilde {\psi }_\nu +
\partial _\nu \tilde {\psi }_\mu } \right) + c \tilde
{\psi }_{\mu \nu } = 0 .
\end{eqnarray}
\end{subequations}
As seen from the comparison between (\ref{eq2.50}) and
(\ref{eq2.2}), these systems are dual
in that one may be derived from the other by the substitutions
\begin{equation}
\label{eq2.51}
\psi _0 \leftrightarrow \tilde {\psi }_0 , \psi
_\mu \leftrightarrow \tilde {\psi }_\mu , \psi
_{\mu \nu } \leftrightarrow \tilde {\psi }_{\mu \nu } .
\end{equation}
Clearly, the use of system (\ref{eq2.50}) with the aim of framing various
gauge-invariant theories on its basis follows the same procedure and gives
the same results as with system (\ref{eq2.2}). So, when in (\ref{eq2.50}) we set $a=0$,
a gauge-invariant theory for a pseudoscalar particle of the
mass $\frac{\sqrt {b c} }{\left| \beta \right|}$ is put forward. But
setting $c = 0$, we arrive at a gauge-invariant theory for a
pseudoscalar particle of the mass $\frac{\sqrt {a b} }{\left| \alpha
\right|}$.
\section{SIMULTANEOUS DESCRIPTION OF MASSLESS FIELDS}
Returning to a set of representations (\ref{eq2.1}) and to tensor system (\ref{eq2.2}), we
consider the case
\begin{equation}
\label{3.1}
b = 0 .
\end{equation}
The following system is obtained:
\begin{subequations}
\label{3.2}
\begin{eqnarray}
\label{3.2a}
\alpha \partial _\mu \psi _\mu + a \psi _0 =
0 ,
\\
\label{3.2b}
\beta ^ * \partial _\nu \psi _{\mu \nu } + \alpha ^ *
\partial _\mu \psi _0 = 0 ,
\\
\label{3.2c}
\beta \left( { - \partial _\mu \psi _\nu + \partial _\nu \psi
_\mu } \right) + c \psi _{\mu \nu } = 0
\end{eqnarray}
\end{subequations}
that in basis (\ref{eq2.3}) is associated with the matrix $\gamma _0 $ of the form
\begin{equation}
\label{3.3}
\gamma _0 = \left( {{\begin{array}{*{20}c}
a \hfill & \hfill & \hfill \\
\hfill & {0_4 } \hfill & \hfill \\
\hfill & \hfill & {c I_4 } \hfill \\
\end{array} }} \right) .
\end{equation}
From system (\ref{3.2}) we get d'Alembert equation (\ref{eq2.12}) for the scalar function
$\psi _0 $ and the second-order equation
\begin{equation}
\label{3.4}
\Box\psi _\mu - \left( {1 - \frac{c \left| \alpha
\right|^2}{a\left| \beta \right|^2}} \right) \partial _\mu \partial
_\nu \psi _\nu = 0
\end{equation}
for the vector $\psi _\mu $. From this it is inferred that we deal with a
massless field. When considering the quantities $\psi _0 $ and $\psi _\mu $
as potentials of this field, we treat equation (\ref{3.2c})
as a definition of the
intensity $\psi _{\mu \nu } $ in terms of the potentials, (\ref{3.2a}) is
additional constraint similar to the Feynman gauge. Then equation (\ref{3.2b})
acts as an equation of motion.
With this treatment, system (\ref{3.2}) and equation (\ref{3.4}) is invariant with
respect to the gauge transformation
\begin{equation}
\label{3.5}
\psi _\mu \to \psi _\mu + \partial _\mu \Lambda ,
\end{equation}
where an arbitrary choice of $\Lambda $ is constrained by (\ref{eq2.15}). Gauge
transformations (\ref{3.5}) and (\ref{eq2.15}) in combination with an additional requirement
(\ref{3.2a}) indicate that, among the four components of the potential $\psi _\mu
$, only two components are independent. They describe a transverse component
of the field under study. One more, longitudinal, component of this field is
described by the scalar function $\psi _0 $. In this way a choice of (\ref{3.1}) in
system (\ref{eq2.2}) leads to a theory of a massless filed with three helicity
values $\pm 1 , 0 .$ This is one of the distinguishing features of
system (\ref{eq2.2}) as opposed to a theory of Duffin--Kemmer for spin 1, that on
a similar passage to the limit results in a massless field with helicities
$\pm 1 .$
Also, note that equation (\ref{3.4}) with due regard for (\ref{3.2a}) may be rewritten
as
\begin{equation}
\label{3.6}
\Box\psi _\mu + \left( {1 - \frac{c \left| \alpha
\right|^2}{a\left| \beta \right|^2}} \right) \frac{a}{\alpha
} \partial _\mu \psi _0 = 0 ,
\end{equation}
from whence it follows that a gradient of the scalar component acts as an
(internal) source of the transverse component of this massless field.
Next we select the case when in system (\ref{eq2.2})
\begin{equation}
\label{3.7}
a = 0 , \quad b = 0 .
\end{equation}
The resultant system
\begin{subequations}
\label{3.8}
\begin{eqnarray}
\label{3.8a}
\partial _\mu \psi _\mu = 0 ,
\\
\label{3.8b}
\beta ^ * \partial _\nu \psi _{\mu \nu } + \alpha ^ *
\partial _\mu \psi _0 = 0 ,
\\
\label{3.8c}
\beta \left( { - \partial _\mu \psi _\nu + \partial _\nu \psi
_\mu } \right) + c \psi _{\mu \nu } = 0
\end{eqnarray}
\end{subequations}
is distinguished from system (\ref{3.2}) by the potential gauge requirement
(compare (\ref{3.2a}) with (\ref{3.8a})).
In this case the matrix $\gamma _0 $ is of the form
\begin{equation}
\label{3.9}
\gamma _0 = \left( {{\begin{array}{*{20}c}
0 \hfill & \hfill & \hfill \\
\hfill & {0_4 } \hfill & \hfill \\
\hfill & \hfill & {c I_6 } \hfill \\
\end{array} }} \right) .
\end{equation}
From (\ref{3.8}) one can obtain equation (\ref{eq2.12}) for the function $\psi _0 $ and
the second-order equation
\begin{equation}
\label{3.10}
\Box\psi _\mu - \frac{\alpha ^ * c}{\left| \beta
\right|^2} \partial _\mu \psi _0 = 0
\end{equation}
for $\psi _\mu $that, similar to system (\ref{3.8}),
is invariant with respect to
gauge transformations (\ref{3.5}), (\ref{eq2.15}).
All this indicates that we deal again
with two interrelated massless fields: vector field with helicity $\pm
1 $ and scalar field with helicity $ 0 $, the gradient of a scalar
field acting as a source of the vector field.
The other two massless analogs of system (\ref{eq2.2}), when
\begin{equation}
\label{3.11}
a = 0 , \quad c = 0
\end{equation}
and
\begin{equation}
\label{3.12}
b = 0 , \quad c = 0 ,
\end{equation}
are associated with the description of a massless field of zero helicity.
Establishing this fact, we will not concern ourselves with the details.
Considering the possibility for simultaneous description of different
massless fields, we next analyze a set of the representations in (\ref{eq2.30}) and
the first-order system of (\ref{eq2.31}).
First, we take the case
\begin{equation}
\label{3.13}
c = 0, \quad a = b .
\end{equation}
In this case system (\ref{eq2.31}) is of the form
\begin{subequations}
\label{3.14}
\begin{eqnarray}
\label{3.14a}
\alpha \partial _\nu \psi _{\mu \nu } + a \psi _\mu =
0,
\\
\label{3.14b}
\beta \partial _\nu \tilde {\psi }_{\mu \nu } + a \tilde
{\psi }_\mu = 0,
\\
\label{3.14c}
\alpha ^ * \left( { - \partial _\mu \psi _\nu + \partial _\nu
\psi _\mu } \right) + \beta ^ * \varepsilon _{\mu \nu \alpha
\beta } \partial _\alpha \tilde {\psi }_\beta = 0,
\end{eqnarray}
\end{subequations}
and the matrix $\gamma _0 $ (\ref{eq2.32}) is transformed to the matrix
\begin{equation}
\label{3.15}
\gamma _0 = \left( {{\begin{array}{*{20}c}
{aI_8 } \hfill & \hfill \\
\hfill & {O_6 } \hfill \\
\end{array} }} \right).
\end{equation}
In (\ref{3.14}) we take components of the tensor $\psi _{\mu \nu } $ as
potentials, assuming the vector $\psi _\mu $ and the pseudovector $\tilde
{\psi }_\mu $ as intensities. Then equations (\ref{3.14a}) and (\ref{3.14b})
are the
intensity definitions in terms of the potentials and (\ref{3.14c}) acts as an
equation of motion.
From system (\ref{3.14}) we derive the second-order equation for the
tensor-potential $\psi _{\mu \nu } $
\begin{equation}
\label{3.16}
\Box\psi _{\mu \nu } = 0.
\end{equation}
Equations (\ref{3.14}) and (\ref{3.16}) are invariant with respect to the gauge
transformations
\begin{equation}
\label{3.17}
\psi _{\mu \nu } \to \psi _{\mu \nu } + \partial
_\mu \Lambda _\nu - \partial _\nu \Lambda _\mu ,
\end{equation}
where an arbitrary choice of the functions $\Lambda _\mu $ is constrained by
\begin{equation}
\label{3.18}
\Box\Lambda _{\mu } - \partial _\mu \partial _\nu \Lambda _\nu
= 0 .
\end{equation}
Equation (\ref{3.16}) and gauge transformations (\ref{3.17}) and (\ref{3.18}) indicate that a
choice of (\ref{3.13}) leads to a theory for a massless particle of zero helicity
carrying spin 1 in the process of interactions.
By the present time, two approaches to the description of such a particle
have been known: (1) Ogievetsky and Polubarinov approach [17] in which an
intensity is represented by the vector (in [17] this particle is called the
notoph) and (2) Kalb-Ramond approach [13], where an intensity is represented
by the antisymmetric third-rank tensor or pseudovector (Kalb-Ramond field).
System (\ref{3.14}) combines the description of both fields in one irreducible
RWE.
In a sense this pattern may be complemented if in (\ref{eq2.31}) we set
\begin{equation}
\label{3.19}
a = 0 , \quad b = 0 .
\end{equation}
As a result, we have the following system:
\begin{subequations}
\label{3.20}
\begin{eqnarray}
\label{3.20a}
\partial _\nu \psi _{\mu \nu } = 0 ,
\\
\label{3.20b}
\partial _\nu \tilde {\psi }_{\mu \nu } = 0 ,
\\
\label{3.20c}
\alpha ^ * \left( { - \partial _\mu \psi _\nu + \partial _\nu
\psi _\mu } \right) + \beta ^ * \varepsilon _{\mu \nu \alpha
\beta } \partial _\alpha \tilde {\psi }_\beta + c \psi _{\mu
\nu } = 0
\end{eqnarray}
\end{subequations}
that is associated with the matrix $\gamma _0 $ of the form
\begin{equation}
\label{3.21}
\gamma _0 = \left( {{\begin{array}{*{20}c}
{O_8 } \hfill & \hfill \\
\hfill & {cI_6 } \hfill \\
\end{array} }} \right) .
\end{equation}
In system (\ref{3.20}) the components $\psi _\mu $ and $\tilde {\psi }_\mu $ are
naturally considered as potentials, and $\psi _{\mu \nu } $ is taken as
an intensity. Then it is invariant with the gauge transformations
\begin{equation}
\label{3.22}
\psi _\mu \to \psi _\mu + \Lambda _\mu
, \tilde {\psi }_\mu \to \tilde {\psi }_\mu
+ \tilde {\Lambda }_\mu ,
\end{equation}
where an arbitrary choice of the gauge functions $\Lambda _\mu , \tilde
{\Lambda }_\mu $ is constrained by
\begin{equation}
\label{3.23}
\alpha ^ * \left( { - \partial _\mu \Lambda _\nu + \partial
_\nu \Lambda _\mu } \right) + \beta ^ * \varepsilon _{\mu \nu
\alpha \beta } \partial _\alpha \tilde {\Lambda }_\beta =
0 .
\end{equation}
In other words, at $\alpha = \beta = 1$ system (\ref{3.20})
represents the well-known two-potential formulation from electrodynamics
(e.g., see [19]) for a massless spin 1 field with helicity $\pm $1.
Thus, a reciprocal complementarity of the theories based on systems (\ref{3.14})
and (\ref{3.20}) is exhibited in their mathematical structure, including that of
the matrix $\gamma _0 $, and also in interpretations of the field components
$\psi _\mu $, $\tilde {\psi }_\mu $, $\psi _{\mu \nu } $ as well as in
properties (helicity) of the particles described.
Of particular interest is the case when in (\ref{eq2.31}) we set
\begin{equation}
\label{3.24}
a = 0 ,\quad c = 0 .
\end{equation}
This results in the system
\begin{subequations}
\label{3.25}
\begin{eqnarray}
\label{3.25a}
\partial _\nu \psi _{\mu \nu } = 0 ,
\\
\label{3.25b}
\beta \partial _\nu \tilde {\psi }_{\mu \nu } + b \tilde
{\psi }_\mu = 0 ,
\\
\label{3.25c}
\alpha ^ * \left( { - \partial _\mu \psi _\nu + \partial _\nu
\psi _\mu } \right) + \beta ^ * \varepsilon _{\mu \nu \alpha
\beta } \partial _\alpha \tilde {\psi }_\beta =
0
\end{eqnarray}
\end{subequations}
and leads to the matrix
\begin{equation}
\label{3.26}
\gamma _0 = \left( {{\begin{array}{*{20}c}
{O_4 } \hfill & \hfill & \hfill \\
\hfill & {b I_4 } \hfill & \hfill \\
\hfill & \hfill & {O_6 } \hfill \\
\end{array} }} \right) .
\end{equation}
For convenience, we rewrite (\ref{3.25}) in the following form:
\begin{subequations}
\label{3.27}
\begin{eqnarray}
\label{3.27a}
\partial _\nu \psi _{\mu \nu } = 0 ,
\\
\label{3.27b}
\beta \left( {\partial _\mu \psi _{\nu \alpha } + \partial
_\alpha \psi _{\mu \nu } + \partial _\nu \psi _{\alpha \mu
} } \right) + b \psi _{\mu \nu \alpha } = 0 ,
\\
\label{3.27c}
\alpha ^ * \left( { - \partial _\nu \psi _\alpha + \partial
_\alpha \psi _\nu } \right) + \beta ^ * \partial _\mu \psi
_{\mu \nu \alpha } = 0 ,
\end{eqnarray}
\end{subequations}
where $\psi _{\mu \nu \alpha } $ is antisymmetric third-rank tensor
dual with respect to the pseudovector $\tilde {\psi }_\mu $.
According to the structure of system (\ref{3.27}), $\psi _\mu $ and $\psi _{\mu
\nu } $ are potentials, $\psi _{\mu \nu \alpha } $ is intensity.
Then equation (\ref{3.27b}) is a definition of the intensity, and (\ref{3.27a})
acts as
an additional constraint imposed on the tensor-potential $\psi _{\mu \nu }
$ and included originally in the system itself. This constraint leaves for
tensor $\psi _{\mu \nu } $ satisfying the second-order equation
\begin{equation}
\label{3.28}
\Box\psi _{\mu \nu } + \frac{\left| \alpha \right|^2}{\left|
\beta \right|^2} \frac{b}{\alpha } \left( {\partial _\mu \psi _\nu
- \partial _\nu \psi _\mu } \right) = 0
\end{equation}
two independent components. As this takes place, system (\ref{3.27}) is invariant
with respect to relative gauge transformations (\ref{3.17}), (\ref{3.18}). Due to an
arbitrary choice of the gauge function $\Lambda _\mu $ constraining by
condition (\ref{3.18}) we have only one independent component for $\psi _{\mu
\nu } $ that is associated with the state of a massless field with zero
helicity.
To elucidate a meaning of the term $ \partial _\mu \psi _\nu -
\partial _\nu \psi _\mu $ in (\ref{3.28}), we turn to the potential $\psi
_{\mu } $. Apart from transformations (\ref{3.17}), (\ref{3.18}), system (\ref{3.27}) is
also invariant with respect to the gauge transformation
\begin{equation}
\label{3.29}
\psi _\mu \to \psi _\mu + \partial _\mu \Lambda
,
\end{equation}
where $\Lambda $ is arbitrary function . From equation (\ref{3.27c}) for $\psi
_{\mu } $ we derive the second-order equation
\begin{equation}
\label{3.30}
\Box\psi _{\mu } - \partial _\mu \partial _\nu \psi _\nu
= 0,
\end{equation}
in combination with (\ref{3.18}) indicating that the potential $\psi _{\mu } $
gives description for the transverse component (helicity $\pm 1$) of the
massless field under study. The expression
\begin{equation}
\label{3.31}
\partial _\mu \psi _\nu - \partial _\nu \psi _\mu
\equiv F_{\mu \nu }
\end{equation}
in equations (\ref{3.27c}) and (\ref{3.28}) may be considered as an intensity associated
with this transverse component. Then equation (\ref{3.27c}) rewritten with regard
to the notation of (\ref{3.31}) as
\begin{equation}
\label{3.32}
\beta ^ * \partial _\mu \psi _{\mu \nu \alpha } - \alpha ^
* F_{\nu \alpha } = 0 ,
\end{equation}
acts as an equation of motion in system (\ref{3.27}).
Thus, a choice (\ref{3.24}) of mass parameters in the initial system (\ref{eq2.31}) leads
to a theory of the generalized massless field with polarizations 0, $\pm $1.
Selection of the parameters
\begin{equation}
\label{3.33}
b = 0 , \quad c = 0 .
\end{equation}
in system (\ref{eq2.31}) also results in a theory of the generalized massless field
with helicities 0, $\pm $1 featuring a dual conjugate of that obtainable in
the case of (\ref{3.24}). Details are beyond the scope of this paper.
\section{MASS GENERATION AND RWE THEORY}
In 1974 in the works [13,14] a mechanism of mass generation was proposed
differing from the well-known Higgs mechanism. Later this mechanism has been
identified as a gauge-invariant field mixing. It's essence is as follows.
Two massless systems of equations are considered cooperatively as initial
systems
\begin{subequations}
\label{4.1}
\begin{eqnarray}
\label{4.1a}
\partial _\nu \psi _{\mu \nu } = 0 ,
\\
\label{4.1b}
- \partial _\mu \varphi _\nu + \partial _\nu \varphi
_\mu + \psi _{\mu \nu } = 0 ,
\end{eqnarray}
\end{subequations}
and
\begin{subequations}
\label{4.2}
\begin{eqnarray}
\label{4.2a}
\partial _\mu \psi _{\mu \nu \alpha } = 0,
\\
\label{4.2b}
- \partial _\mu \varphi _{\nu \rho } - \partial _\nu \varphi _{\rho \mu } -
\partial _\rho \varphi _{\mu \nu } + \psi _{\mu \nu \rho } = 0,
\end{eqnarray}
\end{subequations}
the first system describing an electromagnetic field and the second one
describing
field of Kalb-Ramond. In (\ref{4.2a}) and (\ref{4.2b})
tensor $\psi _{\mu \nu \alpha }
$ is considered to be an intensity. Then into the Lagrangian of this system
an additional term is included
\begin{equation}
\label{4.3}
L_{int} = m\varphi _\mu \partial _\nu \varphi _{\mu \nu }
\end{equation}
without violation of the gauge-invariance for the initial Lagrangian $L_0 $.
This term may be formally treated as an interaction of the fields under
study (so-called topological interaction). Varying the Lagrangian $L = L_0 +
L_{int} $ and introducing the pseudovector $\tilde {\psi }_\mu =
\frac{1}{3!}\varepsilon _{\mu \nu \alpha \beta } \psi _{\nu \alpha \beta } $,
we have a system
\begin{subequations}
\label{4.4}
\begin{eqnarray}
\label{4.4a}
\partial _\nu \psi _{\mu \nu } + m\tilde {\psi }_\mu = 0,
\\
\label{4.4b}
- \partial _\mu \tilde {\psi }_\nu + \partial _\nu \tilde {\psi }_\mu +
m\psi _{\mu \nu } = 0,
\\
\label{4.4c}
\partial _\nu \tilde {\varphi }_{\mu \nu } + \tilde {\psi }_\mu = 0,
\\
\label{4.4d}
- \partial _\mu \varphi _\nu + \partial _\nu \varphi _\mu + \psi _{\mu \nu
} = 0,
\end{eqnarray}
\end{subequations}
where
\begin{equation}
\label{4.5}
\tilde {\varphi }_{\mu \nu } = \frac{1}{2!}\varepsilon _{\mu \nu \alpha
\beta } \varphi _{\alpha \beta } .
\end{equation}
Now in system (\ref{4.4}) we replace $\varphi _\mu $ and $\tilde {\varphi }_{\mu
\nu } $ by the quantities $\tilde {G}_\mu $ and $G_{\mu \nu } $ using the
formulae
\begin{subequations}
\label{4.6}
\begin{eqnarray}
\label{4.6a}
\tilde {G}_\mu = \varphi _\mu - \frac{1}{m}\tilde {\psi }_\mu ,
\\
\label{4.6b}
G_{\mu \nu } = \tilde {\varphi }_{\mu \nu } - \frac{1}{m}\psi _{\mu \nu } .
\end{eqnarray}
\end{subequations}
Finally, system (\ref{4.4}) is reduced to the following form:
\begin{subequations}
\label{4.7}
\begin{eqnarray}
\label{4.7a}
\partial _\nu \psi _{\mu \nu } + m\tilde {\psi }_\mu = 0,
\\
\label{4.7b}
- \partial _\mu \tilde {\psi }_\nu + \partial _\nu \tilde {\psi }_\mu +
m\psi _{\mu \nu } = 0,
\\
\label{4.7c}
\partial _\nu G_{\mu \nu } = 0,
\\
\label{4.7d}
- \partial _\mu \tilde {G}_\nu + \partial _\nu \tilde {G}_\mu = 0.
\end{eqnarray}
\end{subequations}
As seen, system (\ref{4.7}) is reducible with respect to the Lorentz group into
subsystems (\ref{4.7a}), (\ref{4.7b}) and (\ref{4.7c}), (\ref{4.7d}).
The first of them describing a
massive spin 1 particle is interpreted in [13] as an interaction transporter
between open strings. Subsystem (\ref{4.7c}), (\ref{4.7d}) gives no description for a
physical field, as it is associated with zero energy density. However, its
presence is necessary to impart to the latter the status of a
gauge-invariant theory.
Using the formalism of generalized RWE, all the above may be interpreted as
follows. Let us consider a set of representations
\begin{equation}
\label{4.8}
\left( {\frac{1}{2},\frac{1}{2}} \right) \oplus \left(
{\frac{1}{2},\frac{1}{2}} \right)^\prime \oplus 2(1,0) \oplus 2(0,1),
\end{equation}
associated with tensor system (\ref{4.1}), (\ref{4.2}).
It is obvious that on the basis of (\ref{4.8})
one can derive RWE (\ref{eq1.1}) with the
matrices
\begin{equation}
\label{4.9}
\gamma _\mu = \left( {{\begin{array}{*{20}c}
{\gamma _\mu ^{DK} } \hfill & \hfill \\
\hfill & {\gamma _\mu ^{DK} } \hfill \\
\end{array} }} \right) ,
\quad
\gamma _0 = \left( {{\begin{array}{*{20}c}
{O_4 } \hfill & \hfill & \hfill & \hfill \\
\hfill & {I{ }_6} \hfill & \hfill & \hfill \\
\hfill & \hfill & {O_6 } \hfill & \hfill \\
\hfill & \hfill & \hfill & {I_4 } \hfill \\
\end{array} }} \right) ,
\end{equation}
where $\gamma _\mu ^{DK} $ are 10-dimensional Duffin-Kemmer matrices.
Introduction into the Lagrangian of a topological term (\ref{4.3}) results in the
changed form of the matrices $\gamma _\mu $ leaving the matrix $\gamma _0
$ unaltered. Substitutions of (\ref{4.6}) are equivalent to the unitary
transformation restoring the form of $\gamma _\mu $ matrix given in (\ref{4.9}).
As this is the case, the matrix $\gamma _0 $ takes the form
\begin{equation}
\label{4.10}
\gamma _0 = \left( {{\begin{array}{*{20}c}
{mI_{10} } \hfill & \hfill \\
\hfill & {O_{10} } \hfill \\
\end{array} }} \right).
\end{equation}
In this way we actually arrive at RWE reducible to the ordinary
Duffin-Kemmer equation for a massive spin 1 particle and at the massless
fermionic limit of this equation. Nontrivial nature of the mass generation
method, from the viewpoint of a theory of RWE, consists in the fact that on
passage from the initial massless field(s) to the massive one neither the
form of $\gamma _\mu $ matrices nor the rank of singular $\gamma _0 $ matrix is
affected, the procedure being reduced to permutation of zero and unity
blocks of this matrix only. In the process the number of degrees of freedom
(that is equal to three) for a field system is invariable; it seems as if
the notoph passes its degree of freedom to the photon, that automatically
leads to a massive spin 1 particle.
\section{DISCUSSION AND CONCLUSIONS}
Based on the examples considered, the following important conclusions can be
drawn.
\underline {Conclusion 1.}
\textit{Generalized RWE (\ref{eq1.1}) with the singular matrix }
$\gamma _0 $
\textit{ can describe not only massless but also massive fields (particles).
Featuring the gauge invariance, these equations just form the class
of massive gauge-invariant theories.}
As demonstrated in Sec.~II using equations
(\ref{eq2.34}) and (\ref{eq2.42}) as an example,
a theory of generalized RWE suggests also a variant of the generalized
description for massive and massless fields based on RWE irreducible with
respect to the Lorentz group. Thus, we arrive at the following conclusion.
\underline {Conclusion 2. }
\textit{RWE of the form (\ref{eq1.1}) with the singular matrix }
$\gamma _0 $\textit{ can describe the fields
involving both massive and massless components.
In this case it is more correct to refer to massive-massless
gauge-invariant theories rather than to the massive ones. }
As demonstrated in Sec.~III, within the scope of RWE (\ref{eq1.1}),
on adequate selection of
the Lorentz group representations in a space of the wave function $\psi $
and interpretation of its components, one can give the description of a
massless field not only with helicity $\pm $1 but also with helicity 0 as
well as simultaneous description of the indicated fields. Generalizing this
result for the case of arbitrary spin $S$, we can make the following
conclusion.
\underline {Conclusion 3.} \textit{A theory of the generalized RWE
with the singular matrix} $\gamma _0 $
\textit{makes it possible to describe not only massless fields with maximal
(for the given set of representations) helicity }$\pm S$
\textit{, but also fields with intermediate helicity values
as well as to offer a simultaneous description of these fields. }
It is clear that, all other things being equal, a character of the field
described by equation (\ref{eq1.1}) with the singular matrix $\gamma _0 $ is
dependent on the form of this matrix. To find when the singular matrix
$\gamma _0 $ leads to massless theories and when it results in massive or
massive-massless gauge-invariant theories, we examine the Lorentz structure
of the ``massive'' term $\gamma _0 \psi $ in the foregoing cases. It is
observed that in the case of (\ref{eq2.3}), (\ref{eq2.10}), (\ref{eq2.11})
associated with a massive
gauge-invariant spin 1 theory the matrix $\gamma _0 $ (\ref{eq2.11}) affecting the
wave function $\psi $ (\ref{eq2.3}) in the expression $\gamma _0 \psi $ retains
(without reducing to zero) the Lorentz covariants $\psi _\mu , \psi
_{\mu \nu } $, on the basis of which an ordinary (of the form (\ref{eq1.2}))
massive spin 1 theory can be framed. But in the case of a massless theory
given by (\ref{eq2.3}), (\ref{3.2}), (\ref{3.3}) the matrix $\gamma _0 $ in the expression
$\gamma _0 \psi $ retains the covariant $ \psi _{\mu \nu } $, on the
basis of which it is impossible to frame RWE of the form (\ref{eq1.2}) for a massive
particle. A similar pattern is characteristic for the remaining cases: in
all the massive (massive-massless) gauge-invariant theories the matrix
$\gamma _0 $ affecting the wave function $\psi $ retains its covariant
components necessary for framing of an ordinary massive spin 1 or 0 theory;
provided the expression $\gamma _0 \psi $ doesn't involve such a necessary
set of covariants, massless theories can be framed only. This leads us to
the fourth conclusion.
\underline {Conclusion 4. }\textit{Should} \textit{the generalized
RWE (\ref{eq1.1}) with the singular matrix } $\gamma _0 $ in the product
$\gamma _0 \psi $\textit{ retain a set of the Lorentz covariants sufficient
to frame an ordinary (with det }$ \gamma
_0 \ne 0$
\textit{) theory of a massive spin S particle,
this RWE may be associated with a massive gauge-invariant spin S theory.
Otherwise, when this requirement is not fulfilled for any S, RWE (\ref{eq1.1})
can describe a massless field only. }
Proceeding from all the afore-said, we arrive at the following important
though obvious conclusion.
\underline {Conclusion 5.}
\textit{To frame both massive (massive-massless)
gauge-invariant spin $S$ theory and massless theory with intermediate helicity
values from $+S$ to $-S$ we need an extended, in comparison with
a minimally necessary for the description of this spin (helicity),
set of the irreducible Lorentz group representations in a space
of the wave function} $\psi $.
In the present work, when considering spin 1, the above-mentioned extension
has been accomplished by the introduction of scalar representation $\left(
{0 , 0} \right)$ into a set of the representations given by (\ref{eq2.1}) and of
pseudoscalar representation $\left( {\frac{1}{2} , \frac{1}{2}}
\right)^\prime $ -- into a set given by (\ref{eq2.30}). Greater potentialities are
offered by the use of the multiple (recurrent) Lorentz group
representations.
|
\section{Introduction\label{sec:intro}}
Orbital degree of freedom is one of the central ingredients in Mott insulator and
correlated metal at vicinity of metal-insulator transition.
This issue has attracted much attention not only from material viewpoint~\cite{book,JPSJ}, such as colossal-magnetoresisive manganite, superconducting iron punictides, but also from basic theoretical aspects, such as SU(4) spin model~\cite{Li}, and quantum information~\cite{Kitaev}.
In spite of recent great effort, fundamental properties in Mott insulator with orbital degree are still under examination. One of the reasons is its highly frustrated character. Since orbital degree implies anisotropy of the electronic wave function, inter-site orbital interaction explicitly depends on a bond direction in a crystal. Thus, interaction energies for all equivalent bonds cannot be minimized simultaneously.
This intrinsic orbital frustration effect even without geometrical frustration gives rise to a macroscopic number of degeneracy in classical orbital configurations~\cite{Feiner,Khaliulline,Ishihara1,Ishihara2}. It is known that
this degeneracy is lifted by thermal and quantum fluctuations, and non-trivial orbital ordered state is realized~\cite{Kubo,Nussinov}.
Multiple-exchange interaction is another candidate to lift the degeneracy.
The ring-exchange (RE) interaction have been studied for a long time in solid $^{3}$He~\cite{Thouless}
and strongly correlated electron systems.
The RE interaction is much weaker than the nearest-neighboring (NN) exchange interaction in usual correlated electron systems.
However, the RE interaction plays crucial roles to understand Raman and neutron scattering experiments in high-Tc cuprates and spin ladder systems~\cite{Roger,Schmidt,Brehmer}.
This interaction is also clue to lift the degeneracy and gives rise to a non-trivial ground state in frustrated magnets~\cite{Bernu,Misguich}. Recently, the RE interaction has been studied from view point of quantum criticality between two competing orders~\cite{Senthil,Low}.
However, the RE interaction in Mott insulator with orbital degeneracy has been left untouched.
In this Letter, we report the RE interaction effect in orbital degenerate systems.
In particular, the spin-less $e_g$ orbital model is mainly examined.
The RE interaction Hamiltonian is derived by the perturbational method. We analyze the Hamiltonian in the classical orbital state at finite temperature
and the quantum state at zero temperature.
It is shown that a remarkably weak RE interaction destroys the staggered orbital order
caused by the order-by-fluctuation mechanism in the NN exchange interaction.
Magnetic octupole (OP), i.e. a complex wave function for the $e_g$ orbitals, is ordered, in contrast to the NN exchange interaction model.
Ordered moments are largely suppressed due to strong competition between the electrical quadrupole (QP) and magnetic OP interactions.
The RE interaction in the spin-orbital coupled system is also discussed.
Let us start from the RE interaction in the spin-less orbital model.
We consider a Mott insulator with the $e_g$ orbital degree where one electron per site occupies one of the two orbitals, $(u,v)=(d_{3z^2-r^2},d_{x^2-y^2})$, in a cubic lattice.
The effective model is derived from the spin-less Hubbard model with the $e_g$ orbitals
defined by
\begin{eqnarray}
{\cal H}=\sum_{\langle ij \rangle_a \gamma \gamma'}
\left ( t_a^{\gamma \gamma'} c_{i \gamma}^\dagger c_{j \gamma'} +H.c. \right )
+U\sum_i n_{i u} n_{i v} ,
\label{eq:hubbard}
\end{eqnarray}
where $c_{i \gamma}$ is an annihilation operator of a spin-less fermion with orbital $\gamma(=u, v)$ at site $i$, $n_{i \gamma}=c^\dagger_{i \gamma}c_{i \gamma}$ is the number operator,
and $\langle ij \rangle_a$ represents the NN $ij$ pair along a direction $a(=x, y, z)$.
Matrix elements of $t_a^{\gamma \gamma'}$ are given by the Slater-Koster parameters. We define $t \equiv t_z^{uu}$.
By using the conventional perturbational calculation up to the fourth order of $t$, the effective orbital Hamiltonian is obtained as
$
{\cal H}_{eff}={\cal H}_2+{\cal H}_4 .
$
The second order term provides the conventional NN exchange interaction given by
\begin{eqnarray}
{\cal H}_{2}=J\sum_{\langle ij \rangle_a} \tau^{a}_i \tau^{a}_{j} .
\label{eq:h2}
\end{eqnarray}
The $e_g$ orbital degree of freedom is described by the pseudo-spin (PS) operator with magnitude of 1/2 defined by ${\bf T}_i=(1/2) \sum_{\gamma \gamma'} c^\dagger_{i \gamma} {\bf \sigma}_{\gamma \gamma'} c_{i \gamma'}$ with the Pauli matrices $\bf \sigma$.
The operator $\tau_{i}^a$ in Eq.~(\ref{eq:h2})
is the bond-depend linear combination of $T^x$ and $T^z$ defined by
$\tau_i^a=\cos(2 \pi n_a/3) T_i^z -\sin(2 \pi n_a/3) T_i^x$
with a number $(n_x, n_y, n_z)=(1,2,3)$.
The eigen wave function of $\tau^a_i$ with the eigen value of $+1/2 (-1/2)$
is $d_{3a^2-r^2}$ $(d_{b^2-c^2})$ where $(a, b, c)$ are the cyclic permutation of $(x,y,z)$.
The exchange constant is $J=2t^2/U$.
It is known that, in ${\cal H}_2$, a macroscopic number of degeneracy exists in the classical orbital configurations due to the intrinsic orbital frustration effect~\cite{Feiner,Khaliulline,Ishihara1,Ishihara2}.
By taking into account thermal and quantum fluctuations, this degeneracy is lifted and the staggered orbital order at $(\pi, \pi, \pi)$ in the Brillouin zone is realized~\cite{Nussinov,Kubo}.
The ordered pattern is $(d_{3a^2-r^2}, d_{b^2-c^2})$.
This orbital interaction in Eq.~(\ref{eq:h2}) is also obtained
by exchange of the Jahn-Teller phonon~\cite{Okamoto}.
\begin{figure}[t]
\begin{center}
\includegraphics[width=1.0\columnwidth,clip]{ring.eps}
\caption{(color online)
Schematic views of (a) the RE interaction, and (b)
three-up one-down type order of the magnetic OP moment.
Tones represent phases of the wave functions.
}
\label{fig:ring}
\end{center}
\end{figure}
The newly derived fourth-order term of the Hamiltonian is explicitly given by
\begin{eqnarray}
{\cal H}_4
&=&K_{NN} \sum_{\langle ij \rangle_a}
\left ( \tau_i^a \tau_j^a-{\bar \tau}_i^a {\bar \tau}_j^a-T_i^y T_j^y \right )
\nonumber \\
&+&K_{NNN} \sum_{\langle ij \rangle_{a}}'
\left (
\tau_i^a \tau_j^a-5 {\bar \tau}_i^a {\bar \tau}_j^a+\frac{1}{2}T_i^y T_j^y
\right )
\nonumber \\
&+&K_{3NN} \sum_{\langle ij \rangle_a}''
\tau_i^a \tau_j^a
+{\cal H}_R ,
\label{eq:h4}
\end{eqnarray}
where we introduce
${\bar \tau}_i^a=\cos(2 \pi n_a/3) T_i^x + \sin(2 \pi n_a/3) T_i^z$.
The first, second and third terms of Eq.~(\ref{eq:h4}) are for the interaction
between the NN $ij$ sites along $a$,
that between the next NN sites which is perpendicular to $a$,
and that between the third NN sites along $a$.
The exchange constants are
$K_{NN}=3t^4/(2U^3) $, $K_{NNN}=3t^4/(4U^3)$, and $K_{3NN}=4t^4/U^3$.
The last term in Eq.~(\ref{eq:h4}) is the RE interaction:
\begin{eqnarray}
{\cal H}_R=K_{R} \sum_{[ijkl]_a}
\frac{1}{2}
\left (
\tau_i^{a +} \tau_j^{a -} \tau_k^{a +} \tau_l^{a -} +H.c.
\right ),
\label{eq:ring}
\end{eqnarray}
where
$\tau^{\pm a}_i=\tau_i^a \pm i (\sqrt{3}/2)T_i^y$
and $K_R=40t^4/U^3$.
A symbol $[i j k l]_a$ represents sites in a plaquette square which is perpendicular to $a$ [see Fig.~\ref{fig:ring}(a)].
An amplitude of the fourth order terms is represented by a parameter $r_R \equiv \sqrt{K_R/(20J)}$.
A realistic value of $r_R$ is 0.1-0.2 for LaMnO$_3$ and 0.4-0.5 for LaNiO$_3$, although we examine the RE interaction effect in a wide range of $r_R$.
Several characteristics in newly obtained ${\cal H}_4$ are listed.
(i) The RE interaction term, ${\cal H}_R$, is dominant in ${\cal H}_4$ due to its large exchange constant $K_R$. (ii) When we consider ${\cal H}_R$ in a plaquette, the classical stable configuration is that one of the four $\tau_i^{a \pm}$'s is positive (negative) and others are negative (positive).
This state, termed three-up one-down (3U1D) configuration,
competes with the staggered orbital configuration favored in ${\cal H}_2$.
(iii) The $y$-component of the orbital PS $T_i^y$ appears. The eigen states of $T_i^y$ is $(d_u \pm i d_{v})/\sqrt{2}$ which breaks the time-reversal symmetry and represents the magnetic OP with A$_{2g}$ symmetry [see Fig.~\ref{fig:ring}(b)].
The $T^y$ operator in ${\cal H}_R$ is caused by the perturabational processes where four plaquette sites are concerned.
This is in contrast to ${\cal H}_2$ where $T_i^x$ and $T_i^z$ implying the electric QP with the $E_g$ symmetry only appears.
The OP order was discussed in doped manganites and $f$-electron systems~\cite{Takahashi,Brink,Maezono,Kuramoto}, but not in a Mott insulator.
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.9\columnwidth,clip]{classical.eps}
\caption{(color online)
Classical phase diagram. Symbols QP(R), QP(R+M) and OP(X)+QP(R+M) represent the staggered QP order,
the canted QP order and a coexistence of the OP and QP orders, respectively.
Data at $(T, r_R)=(0,0)$ is obtained by the mean-field method.
Inset shows degenerate lines at $(T, r_R)=(0, 0)$ and momenta for the QP orders in the Brillouin zone.
}
\label{fig:classical}
\end{center}
\end{figure}
First we show the RE interaction effect in the classical orbital state.
In Fig.~\ref{fig:classical}, the phase diagram obtained by the Monte-Carlo (MC) simulation is presented as a function of $r_R$.
A three-dimensional $10^3$-site cluster with periodic boundary condition is used.
We adopt the Wang-Landau algorithm where $5 \times 10^{7}$ MC steps is used for standard measurements.
The phase diagram is obtained by the correlation functions and the specific heat.
All fourth-order terms are considered, although the calculation in ${\cal H}_2+{\cal H}_R$ qualitatively reproduces the results in the figure.
As mentioned previously, a macroscopic number of configurations are degenerate at $(r_R, T)=(0,0)$.
In finite temperatures at $r_R=0$, the staggered QP order
at a momentum $(\pi, \pi, \pi)$ is realized.
The RE interaction lifts the degeneracy at the point $(r_R, T)=(0, 0)$ and stabilizes
the canted QP order at $(\pi, \pi, \pi)$ and $(\pi, \pi, 0)$, and other equivalent momenta.
This order is one of the degenerate states at $(r_R, T)=(0,0)$ and competes with the staggered QP order stabilized by the order-by-fluctuation mechanism in ${\cal H}_2$.
It is worth to note that the infinitesimal RE interaction is relevant for the PS configuration, since $(r_R, T)=(0, 0)$ is the degenerate point.
By further increasing the RE interaction, the OP order appears.
In this phase,
the $T^y$ order at $(0, 0, \pi)$ and the non-collinear $T^x-T^z$ order mentioned above coexist.
A mechanism of this coexistence phase is as follows.
As mentioned previously,
favorable QP ordered patterns in ${\cal H}_J$ and ${\cal H}_R$, i.e. the staggered and 3U1D configurations, respectively, compete with each other. This frustration is released by lifting the PS vector from the $T^x-T^z$ plane toward the $T^y$ axis.
The cross term $\tau_i^{a} \tau_j^{a} T_k^{y} T_l^{y}$ in ${\cal H}_R$ determines the ordered pattern; in a plaquette, a parallel alignment of $T^y_i T^y_j$ associated with an antiparallel one of $\tau^a_k \tau^a_l$, and vice versa, is stabilized.
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.7\columnwidth,clip]{bethe.eps}
\caption{(color online)
(a) Squares of the QP moment, and (b) those in the OP moment obtained by the extended Bethe approximation at representative momenta.
Data at equivalent momentum in the Brillouin zone are summed.
}
\label{fig:bethe}
\end{center}
\end{figure}
Next, we introduce the RE interaction effects in the quantum orbital state at $T=0$.
The QP and OP ordered moments are obtained by the extended Bethe approximation.
These are defined by
$M_Q({\bf q})=[|m^{x}({\bf q})|^2+|m^{z}({\bf q})|^2]^{1/2}$
and
$M_O({\bf q})=|m^{y}({\bf q})|$,
respectively, where
$m^\alpha({\bf q})=(1/N)\sum_i \langle T_i^\alpha \rangle e^{{\bf q} \cdot {\bf r}_i}$.
We solve PS states in a $2^3$-site cluster with the mean fields which are obtained self-consistently with the PS states in a cluster.
The RE interaction in ${\cal H}_R$ is decoupled as
$\langle \tau_i^{a +} \rangle \langle \tau_j^{a -} \rangle \langle \tau_k^{a +} \rangle \tau_l^{a -}$ at corner sites and as
$\langle \tau_i^{a +} \tau_j^{a -} \rangle \tau_k^{a +} \tau_l^{a -}$ in NN bonds in a cluster. We numerically confirmed that this method reproduces the results by the lowest order of the hierarchical mean-field approach proposed in Ref.~\cite{Ortitz}.
We also apply this method to the spin model consisting of
the NN exchange and 4-spin RE interactions, and qualitatively reproduce the phase diagram by the exact-diagonalization method in Ref.~\cite{Lauchli} except for the spin nematic phase.
As shown in Fig.~\ref{fig:bethe},
the staggered QP order is stabilized up to a finite value of $r_R$,
in contrast to the classical phase diagram where the transition occurs at $r_R=0$.
This is caused by the quantum order-by-fluctuation mechanism where zero-point fluctuation stabilizes the staggered QP order.
It is worthnoting that the remarkably weak RE interaction ($r_R=0.02$) destroys the staggered QP order and stabilizes the canted QP order favored in the RE interaction.
With increasing $r_R$, the OP correlation function associated with the QP one
appears around $r_R=0.22$.
This phase corresponds to the QP-OP coexistence phase seen in Fig.~\ref{fig:classical}.
One noticeable point is that both the QP and OP correlation functions vanish at about $r_R=0.35$. This result suggests a suppression of the conventional orders due to quantum fluctuation.
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.7\columnwidth,clip]{eta.eps}
\caption{(color online)
(a) Energies and (b) ordered moments in the modified RE interaction Hamiltonian where $\tau_i^{\pm a}$ is replaced by $\tau_i^{\pm a}(\eta)$.
Circles, triangles and bold lines are for the data obtained by the Bethe method, the spin wave approximation and the mean-field approximation.
Vertical broken lines indicate $\eta=3/4$.
}
\label{fig:eta}
\end{center}
\end{figure}
To clarify the last point furthermore, we study the RE term alone
and focus on competition between the QP and OP interactions.
We modify the operator $\tau_i^{a \pm}$ in Eq.~(\ref{eq:ring}) by introducing a parameter $\eta$ as $\tau_i^{\pm a} \rightarrow
\tau_i^{\pm a}(\eta) \equiv
\tau_i^a \pm i \sqrt{\eta}T_i^y$.
It is obvious $\tau_i^{\pm a}(\eta=3/4)=\tau_i^{\pm a}$.
The energy and the ordered moments, defined by $(\sum_{\alpha q} |m^{\alpha}({\bf q})|^2)^{1/2}$, as a function of $\eta$ are shown in Fig.~\ref{fig:eta}.
We adopt the extended-Bethe method and the spin-wave approximation where effects of the zero-point vibration is included.
In the small and large limits of $\eta$, the 3U1D type orders for the QP and OP moments realize, respectively.
In the results by the spin-wave approximation,
a point of $\eta=3/4$ is close to the phase boundary between the QP and OP ordered phases and reduction of the ordered moments occur due to competition between the two.
In the results by the extended Bethe method, this point is inside of the para-orbital phase.
The wave function in this para-orbital phase is given by the linear combination
of the 3U1D state.
We conclude that the suppression of the ordered moments is due to the competition between the OP and QP interactions.
Higher order calculations are required to judge whether the para-orbital phase survives or not.
\begin{figure}[t]
\begin{center}
\includegraphics[width=\columnwidth,clip]{map.eps}
\caption{(color online)
A contour map of the two-body correlation function
$K_Q(\pi,\pi,\pi)$ in (a), that of $K_Q(\pi,\pi,\pi)$ in (b),
and that of a sum of the QP plaquette correlation functions $P_Q$ and the OP plaquette correlation function $P_O$ in (c).
Exact diagonalization method in the $2 \times 2 \times 4$ cluster is adopted.
(d) Phase diagram obtained by the extended Bethe method.
Vertical broken lines indicate $\eta=3/4$.
}
\label{fig:ed}
\end{center}
\end{figure}
Results obtained by the exact diagonalization method helps us to understand whole feature of the phase diagram.
In Fig.~\ref{fig:ed}, we plot contour maps of the two-body correlation functions $K_Q(\pi,\pi,\pi)$, $K_Q(\pi,\pi,0)$,
and a sum of the QP plaquette correlation function $P_Q[=(P^x+P^z)/2]$ and
the OP correlation function $P_O(=P^y)$ in the $r_R-\eta$ plane.
We introduce
$K_Q({\bf q})=(4/N^2) \sum_{ij }
\langle T_i^x T_j^x+T_i^z T_j^z \rangle
e^{{\bf q} \cdot ({\bf r}_i-{\bf r}_j)}$ and
$P^\alpha=(6N)^{-1}\sum_{[ijkl]}(1-16\langle T_i^\alpha T_j^\alpha T_k^\alpha T_l^\alpha\rangle)$.
In a region of small $r_R$, the QP orders at $(\pi,\pi,\pi)$ and $(\pi,\pi,0)$ appear.
On the other side, in large $r_R$,
large plaquette correlation functions in the regions of small and large $\eta$
imply the 3U1D-type QP and OP orders, respectively.
Among them, there is a wide region where magnitudes of all types of the correlation functions are less than 50$\%$ of their maximum values.
This region corresponds to the para-orbital phase obtained by the extended Bethe method
[see Fig.~\ref{fig:ed}(d)].
One noticeable point is that region of the para-orbital phase in the Bethe method increases with decreasing $r_R$ from $r_R=\infty$.
That is, competition between ${\cal H}_2$ and ${\cal H}_R$, as well as the OP and QP interactions in ${\cal H}_R$, contributes to this suppression of the correlation functions.
Finally, we briefly touch the RE interaction in the spin-orbital coupled system. We derive the RE Hamiltonian from the $e_g$-orbital Hubbard model with spin degree of freedom where the intra-orbital Coulomb interaction ($U_0$), the inter-orbital one ($U'$), the Hund coupling $(J_H)$ and the pair-hopping ($J_P$) are taken into account.
Since the full spin-orbital effective Hamiltonian up to the fourth order of the transfer integral is much complicated, we briefly introduce the following two cases: (i) the transfer integrals are diagonal, i.e. $t^{\gamma \gamma'}_a=-\delta_{\gamma \gamma'} t$, and (ii) in the intermediate states of the $e_g^2$ configuration in the perturbational processes, the lowest energy state is only considered.
In the case (i),
the model at $U_0=U'$ and $J_H=J_P=0$ is known to be the SU(4) point where spin and orbital degrees are equivalent.
In spite that the transfer integral is diagonal,
the anisotropic spin order, i.e. the A-type AFM order, is realized between the SU(4) four sublattice phase, favored in the RE interaction, and the ferromagnetic and staggered orbital ordered phase, favored by the Hund coupling.
In the case (ii), the RE interaction includes a new term which is proportional to ${\bf S}_i \cdot {\bf S}_j \times {\bf S}_k T^y_l$ in a plaquette.
This term predicts a coexistence of a chiral spin state and a magnetic OP order.
Summarizing, we have provided a new insight for the orbital physics in correlated electron system beyond the conventional NN exchange interaction, i.e.
the orbital RE interaction.
Effects of the RE interaction in the spin-less $e_g$ orbital system are summarized;
(i)
a remarkably weak RE interaction destroys the staggered QP order stabilized in ${\cal H}_2$,
(ii) the magnetic OP order is realized,
and (iii) large suppression of the ordered moment is caused by the competition of QP and OP
interactions in ${\cal H}_R$, and that of ${\cal H}_2$ and ${\cal H}_R$.
As for (iii), we have a theoretical evidence by the extended Bethe approximation, the spin-wave approximation and the exact diagonalization method that the conventional orders are suppressed by introducing the RE interaction.
However, the issue whether the para-orbital state survives in the higher order calculations or not is left for a future work.
We believe that the present theory does not only serve a new aspect of the orbital physics and the RE interaction, but also help to understand experiments in Mott insulators near the metal-insulator transition.
Authors would like to thank M. Matsumoto, Z. Nussinov, and G. Ortitz for their valuable discussions. This work was supported by JSPS KAKENHI, TOKUTEI from MEXT, CREST, and Grand challenges in next-generation integrated nanoscience. JN is supported by the global COE program of MEXT Japan.
|
\section{INTRODUCTION}
Ultraluminous infrared galaxies (ULIRGs; $L_{IR}>10^{12}\rm L_{\odot}$) were first
discovered in a large number by {\it Infrared Astronomical Satellite} (IRAS)
two decades ago (Rowan-Robinson et al. 1984; Soifer et al. 1984). These objects are among the most
luminous sources in the universe, with most of their energy radiated in the infrared.
Although the powering sources of ULIRGs remain uncertain, there is evidence that
both AGNs and dusty starbursts contribute to their high bolometric
luminosities (see Sanders \& Mirabel 1996; Lonsdale, Farrah, \& Smith 2006 for reviews). The majority of local ULIRGs
are mergers with tidal features commonly seen (e.g. Sanders et al. 1988; Melnick \& Mirabel 1990; Murphy et al. 1996),
and distant ULIRGs are the possible progenitors of today's massive ellipticals (e.g. Barnes \& Hernquist 1996).
Recent large optical photometric surveys show that the color distribution of galaxies are bimodal
(Strateva et al. 2001; Hogg et al. 2003;
Blanton et al. 2003c; Bell et al. 2004a,b). Color-magnitude diagram reveals a red sequence of early type galaxies
with little star formation activity, and a blue cloud of late type
galaxies that form stars. The characteristic number density in the standard luminosity function of
Schechter (1976), ${\phi_{*}}$, has doubled for the red sequence galaxies since $z\sim1$,
but hardly changes for the blue cloud, implying that the number and stellar mass of the red sequence
galaxies have been gradually built up, whereas those of the blue cloud galaxies remain nearly constant
(e.g., Bell et al. 2004b; Faber et al. 2007).
One scenario of building up the most massive galaxies in the
red sequence is major merging of disk galaxies (Bell et al. 2004b), and thus galaxies in the
``green valley'', the region between the red sequence and blue cloud in the
color-magnitude diagram, may represent this transitional type. As ULIRGs are proposed
to be disk-disk mergers that will finally evolve to massive ellipticals
studying the color-magnitude relation of ULIRGs may shed light on the evolutionary scenarios of
building up massive ellipticals.
Galaxy morphologies indicate the distribution of baryonic matter in the galaxies, and are therefore
tracers of its formation and evolution.
The majority of local ULIRGs are merging systems in which disturbed morphologies such as tidal features are commonly observed.
Recent deep sky surveys have generated enormous amounts of galaxy imaging data, and automated morphology
measurement methods are required to analyze those large datasets with much higher efficiencies
than visual inspections provide. There are two quantitative approaches to describing galaxy morphology:
parametric methods and non-parametric methods. Parametric methods fit the light distribution
of a galaxy to pre-defined formulae, such as Sersic index fitting (Blanton et al. 2003c) and
bulge-to-disk ratio (Peng et al. 2002; Simard et al. 2002), and are usually inadequate
to describe the morphologies of complex systems. Non-parametric methods,
such as the concentration, asymmetry, and clumpiness (CAS) system
(e.g. Isserstedt \& Schindler 1986; Abraham et al. 1994; Wu et al. 1999;
Conselice et al. 2000; Conselice 2003), do not rely on pre-defined functions,
and may be the better methods for describing complex systems. Abraham et al. (2003) introduced the
$Gini$ coefficient ($G$), which describes the relative intensity distributions of a galaxy. The $Gini$ coefficient
is correlated with concentration and increases with the fraction of light in compact components.
Lotz, Primack, \& Madau (2004; hereafter LPM04) introduced $M_{20}$, which is the second moment of the brightest 20\% of galaxy
flux. $M_{20}$ is sensitive to spatial distributions of off-axis clumps. $G$ and $M_{20}$ thus should be
sensitive to the presence of merger features and indeed LPM04 found that local ULIRGs are well separated from
normal galaxies in $G-M_{20}$ space, in the sense that ULIRGs have higher $G$ and $M_{20}$ values than normal galaxies.
In this paper we present the results of color-magnitude analysis and $G$ and $M_{20}$ morphological analysis
of 54 ULIRGs from the $IRAS$ 1Jy ULIRG sample (Kim 1995; Kim \& Sanders, 1998) that were observed by Data Release 5
of Sloan Digital Sky Survey. The goals of this paper are to place ULIRGs
in the context of normal galaxies by comparing their color-magnitude relations
and morphologies with those of optically-selected SDSS galaxies, and to construct a low-redshift comparison sample
for studying high-redshift dusty starbursts,
such as Spitzer-identified ULIRGs and submillimeter galaxies (SMGs).
The $IRAS$ 1Jy ULIRG sample is a complete flux-limited sample from the IRAS Faint Source Catalog
(Moshir et el. 1990)
with $S_{60\mu m}>1 \rm Jy$ and $\delta>-40^{\circ}, |b|>30^{\circ}$ (Kim \& Sanders 1998).
The sample contains 118 sources with a redshift range between 0.018
and 0.268 and a median redshift of 0.145. The infrared luminosities range between $10^{12.00}$ and $10^{12.90}\rm L_{\odot}$,
with a median of $10^{12.19}\rm L_{\odot}$.
They are the most IR luminous galaxies in the low redshift universe, and their uniform selection and completeness
make them ideal for a statistical analysis.
Almost all ULIRGs in the sample show visual signs of interactions or mergers, but most are at advanced stages,
and the median R-band luminosity is $\sim2L_*$ (Kim, Veilleux, \& Sanders 2002, hereafter KVS02;
Veilleux, Kim, \& Sanders 2002, hereafter VKS02).
Spectroscopic study showed that the Seyfert fraction increases with infrared luminosity, and $\sim50\%$ of the galaxies with
$L_{IR}>10^{12.3}L_{\odot}$ present Seyfert characteristics. These AGN-powered ULIRGs have
bolometric luminosities and near-IR spectra reminiscent of the local quasars. Another $30\%$ of the whole sample are classified as
H{\sc ii}-region and $\sim40\%$ are classified as LINERs, in both of which the energy sources are thought to be
stellar origin from recent starbursts ($\leq$ few times $10^7$ yr) or shocks
(Veilleux, Kim \& Sanders 1999, hereafter VKS99; Kim, Veilleux, and Sanders 1998)
Near-IR study using 2MASS showed that Seyfert galaxies among the 1Jy sample have much redder colors
and steeper spectral indices in the near-infrared than the rest,
which indicates powerful AGNs as the energy sources (Chen \& Zhang 2006).
Recent $^{12}$CO ($J=1\rightarrow 0$) observations of 17 ULIRGs in the 1Jy Sample, along with 12 other local ULIRGs
showed that their large IR luminosity and gas mass to FIR luminosity ratios are consistent with a model where most of their
luminosity is powered by a brief surge in star formation rate associated with the rapid gas inflow during the merger phase,
although there is also evidence for powerful AGNs with extreme $L_{FIR}/L_{CO}^{\prime}$ among a subset and they are possibly
transitioning to a quasar phase (Chung et al. 2009).
In general, previous studies suggest an evolutionary sequence
of ULIRGs in which ULIRGs start with cool infrared colors ($f_{25\mu m}/f_{60\mu m}<0.2$),
go through a warm ULIRG state ($f_{25\mu m}/f_{60\mu m}>0.2$), and finally evolve to quasars (Sanders 1988; VKS02).
In \S\ref{sample_selection} we present the sample selection. The image analysis is presented
in \S\ref{image_analysis}. We present the color magnitude relation of the ULIRGs in \S\ref{cmr},
their morphological analysis in \S\ref{morphology}, and the summary in \S\ref{summary}.
We use a $\Lambda$CDM cosmology with $H_0=70\rm kms^{-1}Mpc^{-1}$, $\Omega_m=0.3$ and $\Omega_{\Lambda}=0.7$
throughout the paper, except for the absolute magnitudes, for which we use $h=1$,
where $h=H_0/100\rm kms^{-1}Mpc^{-1}$, for a direct comparison with existing studies.
\section{SAMPLE SELECTION}
\label{sample_selection}
The sample consists of all the sources from the $IRAS$ 1Jy ULIRG sample
(Kim 1995; Kim \& Sanders 1998) that have been imaged in the fifth release
of the Sloan Digital Sky Survey (SDSS DR5; Adelman-McCarthy et al. 2007).
The SDSS is the largest homogeneous, wide-area multiwavelength sky survey available
for studying galaxy morphology and color-magnitude distributions in the low redshift universe.
We downloaded SDSS DR5 images at the coordinate position of each source in the 1Jy ULIRG sample,
and did a one-by-one comparison with published optical and near-IR images
(KVS02). This yielded a final sample of 54 ULIRGs identified in SDSS.
The 54 sources span a redshift range between 0.018 to 0.265, with a median of 0.151, and an infrared
luminosity range between $10^{12.00}$ and $10^{12.76}\rm L_{\odot}$ with a median of $10^{12.23}\rm L_{\odot}$.
The mean values of our sample are very close to those in the whole 1Jy sample. We perform
Kolmogorov-Smirnov (K-S) tests between our sample and the whole 1Jy sample for the distributions of redshift,
infrared luminosity, and infrared colors ($F_{25\mu m}/F_{60\mu m}$), and the probabilities that the two
samples are drawn from the same distribution are 80\%, 97\%, and 99\%, for redshift, infrared luminosity,
and infrared colors, respectively. Thus we conclude that our sample is a good representative subset of the whole
1Jy sample.
The redshifts and the FIR luminosities of the sample are presented in the second and third column in
Table \ref{tab:sample}, respectively. There are 14 ULIRGs harboring one or more AGNs
according to optical and near-IR spectroscopic study (VKS99 and references therein),
and they are identified accordingly in Table \ref{tab:sample}. The AGN sample includes all galaxies with
Seyfert activity but excludes all LINERs, which are thought to be low-energy AGNs (e.g. Krolik 1998).
Earlier studies of the 1Jy ULIRG sample concluded that the AGN fraction increases with IR luminosity (VKS99; VKS02).
The comparison sample we choose for studying the color-magnitude relation is the 436,762 galaxies
in the DR4 of the NYU Value Added Galaxy Catalog (NYU-VAGC, DR4; Blanton et al. 2005) with spectroscopic redshifts in the same redshift range
of our ULIRGs. For studying the morphological parameters, the comparison sample we selected consists of the galaxies that
appear on the same SDSS images of the ULIRGs. The size of each image is $13.51\arcmin \times 8.98 \arcmin$,
and there are 1215, 2059, 2019, and 484 comparison galaxies, in the g, r, i, and z band, respectively.
We will discuss the selection method in more detail in \S\ref{morph.calc}.
\section{IMAGE ANALYSIS}
\label{image_analysis}
We obtained pipeline-reduced clean images ({\it fpC*.fit}) from the SDSS DR5 database.
We adopted the skylevels in the observation auxiliary files ({\it tsField*.fit}) and subtracted them from the images.
Each image was checked to ensure the soft skyvalue in the header is correct; for incorrect sky values, we
estimated the sky levels based on the statistics of multiple small empty-sky regions on the same image.
A mosaic of RGB color images of all 54 sources
is shown in Fig. \ref{fig1}. (Refer to the electronic version for the color image).
Each image is 100 pixels across, or $\sim40\arcsec$ at SDSS's pixel scale,
and centered on the source.
The color images were prepared using the method described by Lupton et al. (2004);
$g-$, $r-$, and $i-$band background subtracted images as blue, green, and
red images, respectively. The images were smoothed with a median window of 2
pixels wide, aligned using $r-$band images, and rotated so that north is up.
The relative scales of red, green, and blue images were set to 1:1:2, in order
to emphasize the short wavelength radiation from young stellar populations.
The first visual impression
is that the majority of the galaxies
show disturbed/irregular morphologies, as well as strong color gradients and discontinuities.
Note that there is a smeared three color source on the image of FSC09539+0857. We checked the SDSS Moving
Object Catalog (MOC, the third release; Ivezic et al. 2002) and found that it is an asteroid with Unique SDSS moving object ID s1d059.
The images are arranged according to the $g-$band $G$ and $M_{20}$ coefficients of each source (see {\S\ref{morphology}}),
so that for each column $G$ increases towards the top, and for each row $M_{20}$ increases towards the left.
The coefficients were calculated using the method described by LPM04 (see \S\ref{morph.calc}) and are printed on each image.
Although the images are arranged only relatively in the $G-M_{20}$ parameter space, a clear trend
is seen: more disturbed and multiple nuclei sources have larger $G$ and $M_{20}$ and are
located toward the upper left corner of the mosaic, and the less disturbed sources, with smoother and more symmetric morphology,
have smaller $G$ and $M_{20}$ and tend to be located toward the lower right corner.
Visual inspection of these images also shows a wide range of
morphologies including many close pairs, tidal tails, and otherwise disturbed profiles, in strong support of previous
studies and the general view of ULIRGs as major mergers of gas-rich disk galaxies.
We computed the magnitude within an elliptical aperture whose semi-major axis
is 2 Petrosian radii (Petrosian 1976) for each source.
The Petrosian radius is defined as the radius of a circular aperture on which
the mean surface brightness is equal to 20\% of the average surface brightness
contained inside the aperture. Using an aperture as large as 2 Petrosian radii
ensures that most low-surface-brightness features are included in the photometry,
and the SDSS Petrosian magnitudes we use to comparison are also obtained within
a 2 Petrosian radius aperture for each source (Blanton et al. 2001; Yasuda et al. 2001).
For each source, we align the five images using
the $r$-band image as the reference. Bilinear interpolation is used during the transformation of each image, and flux
is estimated to be conserved within the 0.05\% level. We run $SExtractor$ (Bertin \& Arnouts 1996) on the $r$-band image
and obtain the aperture whose semi-major axis is 2 Petrosian radii, and perform the photometry for all the five bands using the same aperture.
Some sources have disconnected multiple components detected by $SExtractor$,
with each component confirmed to be a part of the ULIRG by its redshift based on previous studies (KVS02).
For these sources, if the components do not have overlapping apertures, we adopt the total flux from multiple
components for photometry; if they have overlapping apertures, we fit the overlapping region with an ellipse and measure
the flux within the ellipse, then subtract it from the total flux to obtain the corrected flux of the entire source.
The airmass and the photometric zero points are obtained from the auxiliary files ({\it tsField*.fit}) and applied when calculating the
magnitudes. The SDSS magnitudes were then converted to AB magnitudes using the AB-SDSS magnitudes corrections.
We find that for the ULIRGs the independently measured apertures are not identical in all five bands.
This is because many ULIRGs are extended and amorphous, and the extensions of low surface brightness
features can be different at different wavelengths. The mean relative difference of the Petrosian radii measured
in $g$ and $r$ band, $\langle{(r_{P,g}-r_{P,r})/r_{P,r}}\rangle $, is $\sim8\%$.
We perform the $g-$band photometry using the apertures measured in the $g$ band,
and find that the measured magnitudes are on average 0.03 mag more luminous than the magnitudes measured
using the $r-$band apertures, and the standard deviation of the difference is 0.10 mag.
Because we use $r-$band apertures to perform the photometry, we adopt this standard deviation as
a typical error of the $^{0.1}g-^{0.1}r$ color for the ULIRGs.
Our main purpose is to locate the positions of the ULIRGs
in the color-magnitude diagram, and compare them with the existing survey data. We
therefore adopt the k-corrections to SDSS bandpasses shifted to $z=0.1$,
the median SDSS galaxy redshift, employing the {\it kcorrect v4.2} package developed by Blanton et al. (2003a).
Each shifted band is denoted by a superscript on the left, e.g., $^{0.1}g$ denotes the SDSS $g$ band shifted to $z=0.1$,
and $^{0.1}g-^{0.1}r$ denotes the k-corrected color between two shifted bands, $^{0.1}g$ and $^{0.1}r$.
Galactic extinction is corrected using the SFD98 dust map (Schlegel, Finkbeiner \& Davis 1998).
We derive the absolute magnitude by applying the distance modulus to the k-corrected apparent magnitude.
We present the photometry results in Table 1.
\section{COLOR-MAGNITUDE RELATION}
\label{cmr}
\subsection{Color and Magnitude of SDSS Comparison Sample}
In Fig. 2 we plot the $^{0.1}g-^{0.1}r$ color against k-corrected $^{0.1}r$-band absolute magnitude $M_{^{0.1}r}$ for all
comparison sample galaxies, which consists of the 436,762 galaxies from NYU SDSS Value Added Catalog (DR4)
within the same redshift range. Individual galaxies are shown as small dots while the contours represent
the galaxy number density. The bimodal distribution of the SDSS comparison galaxies is clearly shown
in the color-magnitude diagram where the red sequence and the blue cloud form two separate groups.
We adopt the empirical definition of the red sequence from Weinmann et al. (2006),
such that the red-sequence galaxies follow the color-magnitude relations,
\begin{equation}
^{0.1}g-^{0.1}r > 0.7 - 0.033(M_{^{0.1}r}-5logh+16.5)
\end{equation}
This relation is shown in Fig. 2 where the red sequence lies above the upper straight line. The centroids of the red sequence and
the blue cloud are separated by $\sim 0.3$ in $^{0.1}g-^{0.1}r$, and between the two populations there is a relatively
low density region called the ``green valley''.
We define the green valley as a strip below the red sequence with a width of 0.1 in $^{0.1}g-^{0.1}r$,
shown as the region between the two straight lines in Fig. 2.
The fiducial width of 0.1 is chosen to be one third of the color
difference between the centroids of the red sequence and the blue cloud, and is also roughly equal to the color difference between the centroid
and the blue limit of the red sequence (the upper straight line). The low density region between the red sequence and the blue cloud
is wider at fainter magnitudes and narrower at brighter magnitudes, and our simple cut may miss some green-valley galaxies
at fainter magnitudes but include more blue-cloud galaxies at brighter magnitudes.
However, this will not affect our main results as we discuss later.
\subsection{ULIRGs Are Optically Bright and Blue}
Table \ref{statistics} lists the statistics of the colors and magnitudes of the ULIRGs and SDSS comparison sample and their subsamples.
We summarize below those optical characteristics of the ULIRGs and their subsamples (AGN and non-AGN) and compare them with
those of the SDSS comparison sample.
The first obvious trend is that the ULIRGs are very luminous in the optical.
The median k-corrected absolute magnitude is $M_{^{0.1}r} = -21.4$, or equivalently 2.5 $L_*$ given $M_{^{0.1}r}=-20.44$
for an $L_*$ galaxy (Blanton 2003b). This is consistent with the median of $R-$band absolute magnitudes for the whole 1 Jy sample
($\sim 2L_*$; KVS02). For comparison the median k-corrected absolute magnitude of the SDSS galaxies
is $M_{^{0.1}r}=-20.4$, which is 1 magnitude fainter than that of the ULIRGs.
The absolute magnitudes of the ULIRGs range from $M_{^{0.1}r} = -19.78$ to $M_{^{0.1}r} = -23.23$
(excluding the two saturated sources), a luminosity range of a factor of 25. Fifty out of the 54 ULIRGs
($\sim93\%$; including the two saturated sources) are more luminous than the median k-corrected $r-$band absolute magnitude of the
SDSS comparison sample. The difference in magnitude distribution between the ULIRGs and the
SDSS comparison sample is further illustrated in Fig. \ref{maghist}.
The ULIRG sample has a much narrower distribution than that of the
comparison sample. The ULIRGs are on average 1 magnitude brighter than that of the entire comparison sample (see panel(a)) as well as
the individual subsample of the red sequence, the green valley, and the blue cloud (panel (c)).
At the highest luminosity bins the ULIRG sample outnumbers all three SDSS sub-samples in fraction
($M_{^{0.1}r}-5logh<-21$; panel (c)).
Another striking trend is that the ULIRGs are very blue, with only a few exceptions.
Their $^{0.1}g-^{0.1}r$ colors span a wide range, from $0.04$ to $1.5$,
and the majority (87\%) of the ULIRGs have optical colors as blue as those of the blue-cloud galaxies.
The median $^{0.1}g-^{0.1}r$ color of the ULIRGs is 0.58 and very close to that of the blue cloud (0.55).
The median $^{0.1}g-^{0.1}r$ color of the red sequence and green valley are 0.95 and 0.78, respectively.
The color distributions of the ULIRGs and the SDSS galaxies are further
illustrated in Fig. \ref{colorhist}(a) and (c).
Considering galaxy color as a function of galalxy luminosity, we limit the $r$-band absolute magnitudes
of the SDSS comparison sample to the same magnitude range as that of the ULIRGs ($-19.8<M_{^{0.1}r}-5logh<-23.2$).
The peak of the ULIRG distribution is $\sim0.3$ mag bluer than that of the SDSS sample within the same $r$-band magnitude range,
but very similar to that of the blue cloud within the same $r$-band magnitude range ($^{0.1}g-^{0.1}r\sim0.6$).
The ULIRGs have a much larger fraction in the bluer bins ($0.5<^{0.1}g-^{0.1}r<0.7$) than the comparison sample,
but a much lower fraction in the redder bins ($0.7<^{0.1}g-^{0.1}r<1.0$), except for the reddest bins
($^{0.1}g-^{0.1}r>1.2$) and the bluest bins ($^{0.1}g-^{0.1}r<0.1$). The ULIRG sample outnumbers the blue-cloud
galaxies in fraction at blue colors ($^{0.1}g-^{0.1}r<0.7$), has a similar fraction compared to the green-valley galaxies
at green colors ($0.7<^{0.1}g-^{0.1}r<0.9$), and has a much lower fraction than the red-sequence galaxies
at red colors ($^{0.1}g-^{0.1}r>0.9$).
The blue colors of the ULIRGs seem inconsistent with the working hypothesis that ULIRGs are dusty, massive systems.
However, the patchy mix of blue and red regions in the optical color images (see
color version of Fig. 1) suggests that the dust extinction is patchy, and the extended distribution of young,
blue starts dominate the overall color of these galaxies.
Veilleux, Sanders \& Kim (1999) have discussed the blue optical continuum colors of the ULIRGs and
suggested that the stellar light suffer less dust extinction than the emission-line gas.
Our ULIRGs would appear even more luminous and bluer
in the optical if corrected for the internal extinction.
The AGN ULIRGs are among the most luminous sources in the ULIRG sample, with a median $M_{^{0.1}r} = -21.9$,
0.8 magnitude more luminous than the median of the non-AGN ULIRGs.
As shown in Fig. \ref{maghist}(b) the AGN ULIRGs are more concentrated
toward high luminosities and outnumber the non-AGN ULIRGs at the highest luminosity bin ($M_{^{0.1}r}-5logh<-23$).
They also have a much flatter color distribution than the non-AGN ULIRGs (see Fig. \ref{colorhist}(b)).
There is little overlap between the AGN ULIRGs and the green-valley galaxies in both magnitude
distribution (see Fig. \ref{maghist}(d)) and color distribution (see Fig. \ref{colorhist}(d)).
\subsection{ULIRGs Are Scarce in the Green Valley}
We overlay the colors and magnitudes of the ULIRG sample on top of the SDSS contours in Fig. 2, with different symbols
representing ULIRGs known to host an AGN (AGN ULIRGs, circles) and H{\sc ii}-region like and LINER galaxies (non-AGN ULIRGs, stars).
The images of FSC12540+5708 and FSC12265+0219 are saturated in the $g$ and $r$ band,
and their magnitudes are shown as upper limits. The error bar shows $\sigma(g_{P,g}-g_{P,r})$,
the standard deviation of difference in $g-$band magnitude measured between $g-$ and $r-$band apertures
(\S\ref{image_analysis}).
The ULIRGs form a distinct group in the color-magnitude diagram: 24 out of the 52 unsaturated ULIRGs ($\sim46\%$)
lie outside the 90\% level contour of the SDSS galaxies in the same redshift range.
Strikingly, only 3 ($6^{+6}_{-4}\%$\footnotemark) lie in the green valley, and none hosts an AGN.
The majority of the ULIRGs (47 out of 54, or $87^{+7}_{-11}\%$) are located below the bottom straight line
that defines the red limit of the blue cloud, and 4 ($7^{+5}_{-6}\%$) are located above the upper
straight line that defines the blue limit of the red sequence.
\footnotetext[1]{The uncertainty on the fraction of sources in any given region of two-dimensional parameter spaces,
e.g. the fraction of red-sequence galaxies in the color-magnitude relations, or the fraction of mergers in the $G-M_{20}$ plot,
is calculated from the possible miscounted number of sources in this region given typical errors in both parameters.}
In the local universe most ULIRGs are major mergers of late-type galaxies, and are proposed to be the precursors of massive
ellipticals (e.g. Toomre \& Toomre 1972; Toomre 1977; Mihos \& Hernquist 1994, 1996; Barnes \& Hernquist 1996).
Given the favorable scenario in which the red sequence is built up by ``wet mergers'' of late-type galaxies
(e.g. Bell et al. 2004, Faber et al. 2007), one may expect a concentration
of merging galaxies in the green valley, the area connecting the blue cloud and the red sequence
in the color-magnitude diagram.
The low fraction (6\%) of our ULIRGs in the green valley, however, suggests that most of these ULIRGs
are not in the transition phase from the blue cloud to the red sequence.
Instead, the majority (87\%) of these ULIRGs have typical blue-cloud colors,
suggesting that they are still undergoing rapid
star formation, and the significant quenching of star formation has yet to start.
AGNs may play a crucial role in quenching star formation and
accelerate the transition of blue-cloud galaxies to the red sequence
(e.g. Hopkins et al. 2006; Croton et al. 2006). Indeed, some studies have found an excess of X-ray selected AGNs
in the green valley (e.g. Georgakakis et al. 2008; Silverman et al. 2008; Hickox et al. 2009; Schawinski et al. 2009). None of the AGN ULIRGs in our sample, however, is located in the green valley, and only
two ($17\%$) are redder than the blue limit of the red sequence.
The remaining 10 ($83\%$) unsaturated sources are bluer than the red limit of the blue cloud.
Hickox et al. (2009) suggested an evolutionary sequence for massive galaxies with halo masses between
$10^{12}\sim10^{13} \rm M_{\odot}$, in which a high accretion rate optical- and infrared-bright SMG/ULIRG/quasar phase
evolves quickly ($\lesssim 10^8\rm yr$), and is followed by a slightly lower accretion rate ``green-valley'' phase
(lasting $\sim$ 1 Gyr) when the objects are detected as X-ray AGNs.
This phase is followed by an intermittent radio AGN phase in which the galaxies reside in the red sequence
with lowest accretion rates. In the context of the Hickox et al. study, the majority of our AGN ULIRGs are still in the early
optical-bright phase and have yet to evolve to the green valley.
\subsection{Color and Magnitude of AGN Host Galaxies}
\label{agn}
In the optical bands the AGN ULIRGs are among the most luminous in the ULIRG sample, and most have
blue colors (see \S\ref{cmr}). AGNs are point-like sources residing in the nucleus region and
typically have blue optical colors.
Thus one may expect the presence of an AGN would affect the color and magnitude of the host galaxy
(e.g. Hickox et al. 2009). However our AGN ULIRGs are low-luminosity Seyfert galaxies and
the AGN emission at optical wavelengths may be highly attenuated.
In order to test whether their high luminosity and blue color is due to
a central AGN we estimate the AGN color and luminosity contribution by a simple model fitting.
\subsubsection{Methodology}
Because the resolution of the images
is limited to only $\sim 1.3\arcsec$ (FWHM, a physical size of $\sim 2.4$ kpc at $z\sim0.1$,
much larger than the physical size of the nuclear region,
we do not try to decompose precisely the central AGN and the host galaxy light distribution from each image.
Instead we subtract a scaled PSF from the nucleus of each galaxy and measure the flux of the scaled PSF and compare
it with the flux of the host galaxy. As there are few unsaturated field stars on each image,
we adopt the PSFs from the SDSS archive. We use two different subtraction methods and obtain
a range of the flux for each subtracted scaled PSF and the host.
In the maximum subtraction method, the maximally scaled PSF is subtracted from the galaxy image
such that the residual is minimized but does not have negative pixels within a PSF FWHM size aperture
around the center. By using this method we obtain
an upper limit to the point source contribution and a lower limit of the host flux.
In the second approach (smooth host subtraction method),
the PSF is scaled and subtracted such that the host has a smooth profile at the center
of the subtraction. The smooth host subtraction is based on the assumption that the galaxy is composed of a central point source
and a smooth host galaxy (e.g. McLeod \& McLeod 2001).
We choose a 3 by 3 pixels area centered on the subtraction center and determine the host to be smooth
if the standard deviation of these nine pixels is less than the average background rms noise on the same image.
For objects with multiple nuclei we subtract the scaled PSF from only the brightest nucleus.
The exceptions are FSC13451+1232 (in $g$ and $r$ band) and FSC15001+1433 (in $g$ band),
where the two nuclei are less than 4$\arcsec$ apart with flux difference less than 20\%.
In these cases we subtracted scaled PSFs from both nuclei. We note that FSC12265+0219 and FSC12540+5708
are saturated in both $g-$ and $r-$band images and are excluded from our analysis.
The flux of the subtracted point source is sensitive to its center position at the sub-pixel level.
To find the center we first fit a two-dimensional gaussian profile for each source, and if the fitted FWHM is less than 30\%
larger than the FWHM of the PSF, we adopt the fitted center as the subtraction center.
Otherwise we adopt the centroid position as the subtraction center.
The centroid is defined as the location where the spatial derivative of the
pixel intensity becomes zero.
The photometric uncertainty of the subtracted point source is calculated assuming it is dominated by the random noise within the central
FWHM size aperture. The mean photometric uncertainties at $g$ and $r$ band are $\sim 0.05\rm\ mag$ and $\sim 0.02\rm\ mag$ for the
maximum subtraction method, and $\sim 0.05\rm\ mag$ and $\sim 0.04\rm\ mag$ for the smooth host subtraction method, respectively.
\subsubsection{AGN Host Galaxies on the Color-magnitude Diagram}
On average the subtracted point sources contribute only a small amount to the total luminosity of
the AGN ULIRGs in both $g$ and $r$ band. This suggests that most of their optical luminosity
is associated with the stellar hosts.
The maximum, median, and mean ratio of the subtracted point source to the
total flux, using the maximum subtraction method, are 67\%, 28\%, 31\% in $g$ band, and 74\%, 27\%, and 31\% in $r$ band,
respectively. Using the smooth host subtraction method, these values are 61\%, 11\%, and 21\% in $g$ band, 74\%, 8\%,
and 21\% in $r$ band, respectively.
The maximum, median, and mean magnitude difference between the hosts and the un-subtracted
sources are 1.19, 0.36, and 0.45 (1.48, 0.34, and 0.47) in $g$ band ($r$ band), using the maximum subtraction method, and
the corresponding values are 1.02, 0.12, and 0.31 (1.03, 0.09, and 0.31) in $g$ band ($r$ band),
respectively, using the smooth host subtraction method.
We overlay the $^{0.1}g-^{0.1}r$ vs. $M_{^{0.1}r}$ color-magnitude diagram for the 12 AGN ULIRGs (filled circles)
and their AGN-subtracted hosts (open circles) on the SDSS color-magnitude contours in Fig. \ref{agnhost}.
Some hosts are much bluer after the subtraction (e.g. FSC11119+3257, bluer by 0.42 using both methods), some are
much redder (e.g. FSC01572+0009, redder by 0.33 using the maximum subtraction method). However,
the median color difference between the host and the original source is only 0.007 and 0.005 mag,
using the maximum subtraction method and the smooth host method, respectively.
Even after the maximum subtraction, only one (FSC11119+3257) source moved closer to the green valley, and the remaining
11 sources are located close to their original positions in the color-magnitude diagram and
barely overlapped with the SDSS comparison galaxies.
This distinction is clearer using the smooth host subtraction method,
after which the hosts are much closer to their un-subtracted counterparts.
In either case most hosts (10 out of 12) remain luminous and blue in the
color-magnitude diagram. With or without the point source subtraction, these AGN host ULIRGs do not appear in the green valley,
the hypothesized transitional stage between the blue cloud and the red sequence that was associated
with X-ray and IR-selected AGNs (e.g. Hickox et al. 2009).
The blue colors of our AGN host galaxies are broadly consistent with other studies of optically-selected AGNs
(e.g. Kauffmann et al. 2003; Jahnke et al. 2004; Silverman et al. 2008).
Kaviraj (2008) studied a local LIRG ($L>10^{11}L_{\odot}$) sample and also found that virtually all LIRGs
are in the blue cloud and the AGN has no impact on the star formation in their host galaxies.
In a broad agreement with out findings, Cowie \& Barger (2008) also found that the bulk of
extinction-corrected 24$\mu$m selected sources in the GOODS-North field lie in the blue cloud.
Our study is limited by the small sample size, and the point source decomposition is limited by the spatial resolution of SDSS.
A more complete sample observed with HST or the next generation James Webb Space Telescope ($JWST$; Gardner et al. 2006 )
is necessary in order to fully characterize the optical properties of the host galaxies in AGN ULIRGs.
\subsection{Color-magnitude Relation: ULIRGs vs Low-$z$ QSOs and Low-$z$ Type 2 Quasars}
Motivated by the high infrared and bolometric luminosity, it has been proposed that
ULIRGs are the early stage of dusty quasars, and a ULIRG-QSO evolutionary scenario has been suggested
(Sanders et al. 1988). Although the sources of the bolometric luminosity for ULIRGs are nearly completely
obscured by dust in the visible bands, a comparison of their host properties with those of QSOs should
provide an important test for the proposed evolutionary scenario. For example, if the ULIRGs and QSOs
are identical except for a geometrical difference such as the viewing angle and dust geometry,
then the optical properties of their hosts should be largely identical. If a purely temporal
evolution and an AGN feedback process removing gas and dust are the key difference
(e.g. Narayanan et al. 2009), then a significant overlap with a systematic shift is expected in the
host color-magnitude relation.
The QSO sample we selected is composed of 1070 objects in the SDSS DR5 quasar catalog
(Schneider et al. 2007) within the same redshift range ($0.018 < z < 0.265$) as the ULIRGs.
The SDSS QSOs are type 1 quasars, selected based on their broadband colors, with $M_i<-22$ mag and
spectroscopically confirmed to have broad lines (FWHM greater than 1000 $\rm kms^{-1}$).
We also selected a type 2 quasar sample for comparison, which consists of 402 objects in the SDSS
type 2 quasar catalog (Reyes et al. 2008) within the same redshift range,
selected based on their optical emission lines.
Systematic photometric difference should be minimal among the SDSS QSOs, type
2 quasars, and the ULIRGs due to the shared origin of the data. For SDSS
QSOs and type 2 quasars, we apply the k-correction and extinction correction
using the same procedure as for the ULIRGs to obtain the absolute magnitudes.
We plot in Fig. \ref{qso} the $^{0.1}g-^{0.1}r$ colors vs $M_{^{0.1}r}$ of the QSOs and the type 2
quasars with dots and contours, on top of the contours of the SDSS field
galaxies and the ULIRG symbols.
Strikingly, the distribution of the type 2 quasars peaks in the green valley where the X-ray
and IR-selected AGNs are found (e.g. Hickox et al. 2009). In comparison, the SDSS QSOs only
overlap slightly with the blue cloud and extend toward very blue colors. The difference in
their distributions in the color-magnitude diagram is at least partly due
to selection effects. The SDSS QSOs are selected unobscured type 1
quasars where the blue continuum and high luminosity of the central
AGNs are visible, while the type 2 quasars are selected obscured AGNs
and thus the dust and gas block the view to the central blue and
luminous AGNs. Studying whether these green type 2 quasars represent
galaxies in a transitional stage from the blue cloud to the red sequence by
AGN quenching, or their green colors are simply color combinations of redden
obscured AGNs and the host galaxies, is beyond the scope of this paper.
The ULIRGs are distributed very differently from the SDSS QSOs and
the type 2 quasars in the color-magnitude diagram. The mean absolute
magnitude of the ULIRGs ($\langle M_{^{0.1}r}\rangle=-21.4$) is very close to that of the SDSS
QSOs ($\langle M_{^{0.1}r}\rangle=-21.3$), but their mean color ($\langle ^{0.1}g-^{0.1}r\rangle =0.57$) is 0.35 mag
redder than that of the SDSS QSOs (0.22). The SDSS QSOs occupy a
narrow range in $M_{^{0.1}r}$, overlap only slightly with the ULIRGs in the color magnitude diagram,
and extend farther to bluer colors than the ULIRGs. In comparison, the $\langle M_{^{0.1}r}\rangle$ of
ULIRGs is 1.1 mag more luminous than that of the type 2 quasar
(-20.3), and their $\langle ^{0.1}g-^{0.1}r\rangle$ is 0.14 mag bluer (0.57 compared to 0.74).
The ULIRGs overlap with the type 2 quasars only at faint magnitudes
($M_{^{0.1}r}>-21$), and all but one AGN ULIRGs are more luminous than 90\% of
the type 2 quasars. We also implement two-dimensional Kolmogorov-Smirnov tests
(KS test, Peacock et al. 1983; Fasano \& Franceschini
1987) between different samples in the color-magnitude two dimensional
parameter space, and the results are listed in Table \ref{KStable}. The test
results indicate that neither the ULIRG sample nor the AGN-ULIRG
subsample is drawn from the same underlying distribution with the SDSS QSO
sample or with the type 2 quasar sample.
\subsection{Evolutionary Tracks on the Color-magnitude Diagram for ULIRGs}
\label{sedmodels}
We model the color-magnitude evolutionary tracks for the ULIRGs using the stellar synthesis model BC03 (Bruzual \& Charlot 2003)
in order to gain some insight into the distribution and evolution scenarios of ULIRGs on the color-magnitude diagram.
Rather than providing a full solution to the SEDs with best-fit parameters, the primary goal of our models is to demonstrate
qualitatively the possible evolutionary paths these galaxies may follow on the color-magnitude diagram
and to establish the associated timescales.
We start with the simplest cases in which the star formation
histories are: 1) a passively evolved single stellar population burst (SSP), and 2) a constant star formation rate (constant).
Fig. \ref{evolvtrack} shows the evolutionary tracks for these two models with a solid line (SSP) and a dashed line (constant).
Both tracks are k-corrected to $z=0.1$ in the same manner as the ULIRGs. The SSP track is normalized
so that it passes through the median color ($^{0.1}g-^{0.1}r = 0.58$) at the median absolute magnitude of the ULIRGs
($M_{^{0.1}r} = -21.43$, or $L_{r}=2.5L_{r,*}$). The constant track is normalized to reach 2.5$L_{*}$ at $t=20$Gyr.
We use solar metallicity in our models for simplicity and ignore intrinsic extinction
as we are primarily concerned with the unobscured stellar population.
Three symbols are plotted on each track to mark three ages, 1 Gyr, 2 Gyr, and 5 Gyr.
In the SSP model, we find that the stellar populations of ULIRGs have ages spanning from
200 Myr to 8 Gyr based on their optical colors, with a median age of $\sim 1$ Gyr.
One exception is the extremely red source FSC11119+3257 whose color is too red to be reproduced
by the SSP model.
Our modeling suggests that a galaxy with a median age takes 1.5 Gyr to reach the red sequence.
In the constant star formation case, the reddest color of the stellar population (at $t=20$Gyr) is $^{0.1}g-^{0.1}r = 0.43$,
which is bluer than 80\% of the ULIRGs. This implies that in the constant star formation case
the star formation has to be quenched in order to move the ULIRGs to the red sequence.
Next we construct a model with a more complex star formation history
and apply different metallicity and dust treatments to try to understand their effects on the color and magnitude.
Our model assumes that: (a) a ULIRG is a merger system consisting of two equal-mass late type galaxies
and that the merger triggers a starburst; (b) the star formation history consists of an SSP component associated
with each progenitor and an exponentially decaying starburst triggered by the merger; (c) the SSP components are identical
with the same age and the starburst is triggered when the SSP components evolve to the median
color and magnitude of the ULIRGs; (d) the starburst forms as much as 10\% of its progenitors' stellar mass
with an e-folding time of $10^7$ yr.
The timescale of $10^{7}$ yr is close to the dynamical timescale of ULIRGs such as Arp 220 ($6\times10^6$ yr;
Mauersberger et al. 1996) and 10\% is a fiducial fraction of mass formed during a merger that is
consistent with the results in the numerical simulations of disk mergers (e.g. Mihos \& Hernquist 1996).
Given that the progenitors are two $L_*$ galaxies with $M_* \sim 10^{11}\rm M_{\odot}$,
the calculated star formation rate starts with a maximum value of $\sim2000\rm M_{\odot}yr^{-1}$
and decreases to several hundred $\rm M_{\odot}yr^{-1}$ after an e-folding time.
Although these star formation rates seem extremely high, they are required by the observed high FIR luminosities.
We calculate the star formation rate of the ULIRGs from their FIR luminosities using the
FIR-SFR relation (Kennicutt 1998). The star formation rates range from 170 $\rm M_{\odot}yr^{-1}$ to
1370 $\rm M_{\odot}yr^{-1}$, with a median of 270 $\rm M_{\odot}yr^{-1}$, which are consistent with our model predicted values.
We examine three different metallicities, $Z=0.008$, $Z=0.02$, and $Z=0.05$,
corresponding to a sub-solar, solar, and super-solar metallicity, respectively,
and construct three base models using this
star formation history.
We assume that the starburst has the same metallicity as the passively evolved progenitor populations.
We set up a grid of different dust extinctions following Charlot \& Fall (2000) for each metallicity,
but apply the dust extinction only to the starburst population.
There are two parameters in the dust treatment method: the total effective V-band optical depth $\tau_{V}$
that affects stars younger than $10^{7}$ yr,
and the fraction $\mu$ of extinction that comes from the diffuse ISM and thus affects stars of all ages.
We choose $\tau_{V}$ between 1 to 10 with a step size of 1.
Bruzual \& Charlot (2003) noted that the average value of $\mu$ is 0.3, and we adopt this value for our modeling.
Our choice of optical depth is conservative, based on the common concept
that these galaxies are dusty starbursts and/or AGNs. However we reiterate that the ULIRGs on average are blue and bright,
and in order to produce the blue colors the dust attenuation
associated with the escaping optical emission has to be modest.
As discussed earlier, it is also possible that the geometry of the dust distribution in the system is patchy
so that the high infrared luminosity originates from the most dust-attenuated regions, while the optical emission
comes from relatively less attenuated regions, e.g. tidal structures at large distances.
We plot nine modeled tracks along with the ULIRGs on the color-magnitude diagram in Fig. \ref{fig8}.
The tracks start from the time when the starburst is induced.
The three rows of panels show sub-solar, solar, and super-solar metallicity,
from top to bottom, respectively. For each metallicity there are three tracks with $\tau_V\ = 1,\ 5,\ 9$,
from left to right, respectively. The tracks are k-corrected to $z=0.1$ and normalized to the median $M_{^{0.1}r}$
which is roughly 2.5 $L_*$. Symbols are plotted on the tracks to mark the age of 1, 2, and 5 Gyr after the starburst starts.
These tracks clearly suggest that both metallicity and dust extinction play important
roles in our understanding of the evolution of these systems.
Dust extinction mostly affects the blue optical colors of the systems younger than several hundred million years.
For all metallicities, the stellar populations reach their bluest colors in less than 20 Myr, but
the bluest color a stellar population can reach becomes increasingly redder with increasing dust extinction.
The blue colors of many ULIRGs can be produced only with the lowest dust extinctions. However, there is almost no color difference
between different dust extinctions when the populations are older than 1 Gyr and their $^{0.1}g-^{0.1}r$ colors
are redder than 0.7, and any systems with the same metallicity will redden to the red sequence at roughly the same time.
Metallicity affects the range of colors and the timescale of the color evolution of the stellar populations,
and higher metallicity systems span wider ranges in the color space, and evolve more quickly.
As shown in Fig. \ref{fig8}, the bluest ULIRGs can be reproduced only by the highest metallicity track.
After 20 Gyr, the super-solar track is 0.25 mag redder than the sub-solar case.
It takes a sub-solar population 2 Gyr to evolve to the red sequence, 1.5 Gyr for the solar population, and only 1 Gyr for the
super-solar one.
In reality the star formation history is almost certainly more complicated than the simple models we discuss here.
In the third model we use a simplified version of star formation history produced from numerical simulations of
gaseous disk mergers by Springel, Di Matteo, and Hernquist (2005). Springel et al. used GADGET-2 (Springel 2005) to trace
star formation in a merger of two spiral galaxies and studied the effects of back hole feedback on the quenching of star formation.
The two spiral galaxies have equal dynamical masses of $3.85\times 10^{12}\rm M_{\odot}$,
and form stars at an exponentially decaying rate before
the merger, with the pre-merger star formation rate $\sim 200\rm M_{\odot}yr^{-1} $. The merger-triggered starburst
happens at $t=1.5\rm Gyr$ and the peak star formation rate is $\sim 2000\rm M_{\odot}yr^{-1}$.
In the black hole feedback scenario the star formation is quenched right after the starburst and the star formation rate
decreases to less than 1 $\rm M_{\odot}yr^{-1}$ at $t=2.5\rm Gyr$, while in the no feedback scenario the star formation rate
decays slowly and remains at several solar masses per year for several Gyr. The intrinsic extinction is ignored in this model.
We plot the evolutionary tracks in Fig. \ref{evolvtrack} with the dotted line representing the track without AGN feedback and the
dashed line representing the one with AGN feedback. Three symbols are plotted on each track to mark the ages of 1, 2, and 5 Gyr.
We find that some of our most luminous ULIRGs are located very close to these tracks in the color-magnitude diagram.
Both tracks evolve more slowly than the SSP track on the same plot because the starburst produces new young blue stars.
Between the two tracks, the one with AGN feedback clearly reddens more quickly than the one without AGN feedback:
it takes 8.5 Gyr (or 7 Gyr after the starburst) for the galaxies without AGN feedback to evolve to the red sequence,
which is 1.5 Gyr longer than the one with AGN feedback. The difference in the timescales
is consistent with the scenario in which AGN feedback quenches the star formation effectively, and consequently helps form
the color bimodality of galaxies.
In summary, these simple models suggest possible evolutionary paths of the ULIRGs on the color-magnitude diagram. We conclude
that constant star forming galaxies cannot evolve to the green valley without the star formation quenched. Galaxies
with lower star formation activity evolve to the red sequence more quickly and spend shorter time in the green valley,
as seen in the SSP model and the numerical simulation model. Dust affects mostly the blue colors of the stellar population,
and has almost no influences on the red colors and the timescale of evolution to the red sequence. Stellar population
with higher metallicity explores a wider range in the color space, and evolves more quickly to the red sequence.
The color evolution of the ULIRGs may be presented by any one of the models with different masses and dust extinctions.
For example, the evolutionary track of the most luminous ULIRGs in our sample may be presented by either the numerical simulation
models with slightly different mass, or by the SSP track in Fig. \ref{evolvtrack} with ten times more mass; the evolutionary
track of a ULIRG with the median color and magnitude may be presented by the numerical simulation models with 90\% of the
optical emission being attenuated. It is also possible that the color evolution of a galaxy is the result of composite
star formation histories, so that the galaxy may leave the blue cloud and enter the red sequence for more than once.
We will not discuss these more complicated scenarios because they are out of the scope of this paper.
\section{MORPHOLOGY}
\label{morphology}
\subsection{$Gini$ and $M_{20}$ Calculations}
\label{morph.calc}
We use the code developed by Lotz et al. (LPM04) to calculate the morphology coefficients $G$ and $M_{20}$
for both the 54 ULIRGs and the comparison sample, galaxies appearing on the same SDSS images
of the ULIRGs. The Gini coefficient, $G$, is defined as
\begin{equation}
G = \frac{1}{\mid{\bar{X}}\mid n(n-1)}\sum_{i}^{n}(2i-n-1)\mid X_i\mid
\end{equation}
where $n$ is the number of pixels and $X_i$'s are the sorted pixel intensities in increasing order. The total second-order moment
$M_{tot}$ of the pixels assigned to the galaxy is defined as
\begin{equation}
M_{tot} = \sum_{i}^{n}f_i[(x_i-x_c)^2+(y_i-y_c)^2],
\end{equation}
where $f_i$ is the flux in each pixel in decreasing order, $x_i$ and $y_i$ are the coordinates of the pixel, and $x_c$, $y_c$ is the center of the
galaxy, computed by minimizing $M_{tot}$. $M_{20}$ is the normalized second-order moment of the brightest 20 percent of the galaxy's
flux, such that
\begin{equation}
M_{20} = {\rm log}_{10}(\frac{\sum_i M_i}{M_{tot}})\ {\rm while}\ \sum_i f_i < 0.2f_{tot},
\end{equation}
where $f_{tot}$ is the total flux of the galaxy.
As noted by LPM04 and Lotz et al. (2008b), the quantity $G$ is very sensitive to the ratio of low surface brightness pixels
to high surface brightness pixels, so a well-defined segmentation map is essential to measure $G$ and $M_{20}$ accurately.
Also, as pointed out by Lisker (2008), measurement uncertainties of $G$ are minimized when
using the Petrosian radius as the aperture size.
We thus adopt the same approach as LPM04 and assign the pixels brighter than the surface brightness at the Petrosian radius
(semi-major axis as measured in an elliptical aperture) to the galaxy.
The measured coefficients are not robust for a source
with the signal-to-noise ratio per pixel within the Petrosian radius ($\langle S/N \rangle $)
less than 2.5, or with a Petrosian
radius ($r_P$) less than twice the size of FWHM of the PSF ($\sim1.3\arcsec$ in SDSS) (LPM04 and Lotz et al. 2008b),
where the signal-to-noise ratio per pixel is defined as the ratio of the mean flux and the mean rms noise within an aperture.
We therefore exclude all sources with $\langle S/N\rangle <2.5$, and/or $r_P<2.6\arcsec$.
The $u-$band images are noisy in general: almost 60\% of the ULIRGs have $\langle $S/N$\rangle $ less than 2.5, and thus we do not
include the $u-$band results further in our analysis. We also exclude possible candidates for stars in the comparison sample by rejecting
sources with the
$SExtractor$ star/galaxy separation parameter CLASS\_STAR greater than 0.9. The final selection yields the measured morphology parameters
for 50, 50, 51, and 42 ULIRGs, and 1215, 2059, 2019, and 484 comparison galaxies, in the $g$, $r$, $i$, and $z$ band, respectively.
\subsection{Morphology of the ULIRGs}
In Fig. \ref{gm20} we plot the $g-$, $r-$, $i-$, and $z-$band $G$ against $M_{20}$ of the comparison galaxies
using contours and dots, where the contours are 10 equally spaced levels between 10\% and 90\% of the peak
number density, and dots show galaxies located outside the lowest 10\% contour level.
On top of them we overlay the morphology parameters of ULIRGs with circles (AGN ULIRGs)
and stars (H{\sc ii}-region like/LINERs).
We also plot the empirical lines that divide the parameter space into four morphology regions:
mergers, elliptical galaxies, irregular galaxies,
and spiral galaxies, according to the classifications for $z\sim0$ galaxies described in LPM04 and Lotz et al. (2008a):
\begin{eqnarray}
\rm{Merger:} & G > -0.115M_{20}+0.384 & \nonumber \\
\rm{Elliptical:} & G < -0.115M_{20}+0.384 &\&\ G > 0.115M_{20}+0.769 \nonumber \\
\rm{Irregular:} & G > 0.115M_{20}+0.697 & \&\ G < 0.115M_{20}+0.769 \\
& \&\ G < -0.115M_{20}+0.384 & \nonumber \\
\rm{Spiral:} &G < -0.115M_{20}+0.384 & \&\ G < 0.115M_{20}+0.697 \nonumber
\end{eqnarray}
The distributions of the ULIRGs at all wavelengths are consistent in that they
form a similarly heterogeneous group in this parameter space.
The centroids of the distribution
in $G-M_{20}$ are [0.55,-1.42], [0.55,-1.43], [0.57,-1.47], [0.54,-1.45], in $g$, $r$, $i$, and $z$ band, respectively,
all of which are located very close ($\Delta G<\pm 0.02$ for any given $M_{20}$) to the separation line
between merger and non-merger galaxies.
There are only 21($42^{+18}_{-24}\%$), 21($42^{+20}_{-22}\%$), 25($49^{+19}_{-27}\%$), and 15($36^{+24}_{-26}\%$) ULIRGs
located in the merger region, in $g$, $r$, $i$, and $z$ band,
respectively. These are very low fractions compared to the $z\sim0$ results where the majority ($80\%$) of ULIRGs are located
in the merger region (LPM04). The distributions of the ULIRGs in the parameter space vary slightly with wavelengths:
the merger fraction is the highest ($49\%$) using the $i-$band derived parameters, and lowest ($36\%$) using the $z-$band derived ones.
Although the morphology parameters of the ULIRGs span a wide range in the parameter space,
their $G-M_{20}$ distribution is clearly different from that of the comparison sample galaxies.
The distribution centroids of the comparison galaxies in the four bands are all located in the empirical region of spiral galaxies,
and only 56($5^{+14}_{-4}\%$), 117($6^{+19}_{-5}\%$), 146($7^{+21}_{-6}\%$), and 26($5^{+21}_{5}\%$) galaxies are found in the empirical region of mergers,
in $g$, $r$, $i$, and $z$ band, respectively.
The poor physical resolution of the SDSS images
(700 pc pixel$^{-1}$ at $z\sim 0.1$) may lead to higher random and systematic uncertainties in $G$ and $M_{20}$
(see LPM04), and thus careful estimation
of the uncertainties in $G$ and $M_{20}$ is necessary.
Using image and noise simulations (described below in \S\ref{morph.discussion}), we estimate the uncertainties of
morphology measurements ($\Delta G\lesssim0.05$ and $M_{20}\lesssim0.3-0.4$)
in SDSS images and plot the largest uncertainties as error bars in Fig. \ref{gm20}.
We find that a galaxy seen against higher background noise tends to move
to the left lower corner in the $G-M_{20}$ plot, and we plot a vector to illustrate the typical changes in
$G$ and $M_{20}$ resulting from adding background noise until the source only barely satisfies the
signal-to-noise ratio and Petrosian radius criteria.
The vector shows that a source selected as a merger at high signal-to-noise ratio could move away from the
merger region and stay in the late-type galaxy region if it is immersed in higher background, and consequently,
the merger fraction selected by $G$ and $M_{20}$ is likely to be a lower limit.
Although less than half of the ULIRGs have empirical merger $G-M_{20}$ relations, the heterogeneous
distributions are qualitatively consistent with what was found in the recent numerical simulations
of merging galaxies of Lotz et al. (2008b). We will discuss these noise simulations further, together with the uncertainties
in measuring the morphologies, in \S\ref{morph.discussion}.
AGN ULIRGs are slightly more concentrated in the merger region than non-AGN ULIRGs: there are 45\%, 50\%, 60\%, 50\% of
AGN ULIRGs in the merger region, in $g$, $r$, $i$, and $z$ band, respectively, compared to 41\%, 40\%, 48\%, 30\%
for non-AGN ULIRGs.
However the difference is not statistically significant given the small number of AGN ULIRGs (11 in $g$ band
and 10 in other bands based on the signal-to-noise ratio and size selection criteria).
There is no significant difference between the mean $G$ ($\langle G\rangle$) and $M_{20}$ ($\langle M_{20}\rangle$)
for the AGN ULIRGs and non-AGN ULIRGs: for AGN ULIRGs, $\langle G\rangle$ is 0.58, 0.57, 0.59, 0.58,
and $\langle M_{20}\rangle$ is -1.48, -1.48, -1.36, -1.48, in $g$, $r$, $i$, and $z$ band, respectively;
in comparison, $\langle G\rangle$ is 0.54, 0.54, 0.56, 0.53, and
$\langle M_{20}\rangle$ is -1.42, -1.45, -1.49, -1.42 for non-AGN ULIRGs in the same bands.
\subsection{Are the ULIRGs Mergers in $G-M_{20}$ space?}
\label{morph.discussion}
Our visual inspection shows that almost all the ULIRGs are disturbed systems (Fig. \ref{fig2}), and this is consistent with
the visual classifications by VKS02, who found that all but one of the 118 $IRAS$ 1Jy sample show visual merger features.
We perform the $G-M_{20}$ analysis and find that more disturbed galaxies do have higher $G$ and $M_{20}$ coefficients.
This is illustrated in Fig. \ref{fig1} where
more disturbed galaxies tend to occupy the upper left corner of the mosaic with higher values of both $G$ and $M_{20}$.
However our studies also show that ULIRGs are a heterogeneous group in $G-M_{20}$ space: only slightly less than half of the
sources lie above the solid line in each panel of Fig. \ref{gm20}, the region where most local mergers and ULIRGs
were found by LPM04. Interestingly, there is only one source located in the early type region,
and all the remaining sources fall in the region where irregular and late-type galaxies are located.
This heterogeneous distribution seems to be inconsistent with the result of LPM04
that 80\% of local ULIRGs are located in the typical merger regions in the parameter space.
However, the spread in the parameter space we found is supported by numerical simulations.
Lotz et al. (2008b) analyzed the morphological parameters in SDSS $g-$band images of mergers of equal mass gas-rich
spirals by using the N-body/hydrodynamic simulation code GADGET (Springel, Yoshida, \& White 2001)
and Monte Carlo radiative transfer code SUNRISE (Jonsson 2006; Jonsson et al. 2006),
and find that the mergers are most disturbed in $G-M_{20}$ at the first pass,
but they can have normal galaxies morphologies at other merger stages.
They also noted that two-thirds of the ULIRGs in LPM04 exhibit double or multiple nuclei,
and therefore are more effectively selected by $G-M_{20}$.
The timescale is only 0.2 $-$ 0.6 Gyr for a merger to appear in the empirical merger region defined by LPM04,
and the range in this time scale depends on dust, orbit parameters, and observing orientations.
Thus at face value, our results imply that about half of our ULIRGs have been captured within this time window.
The fact that the other
half of our ULIRGs do not appear in the empirical merger region does not mean that they are not mergers. Instead, they might be in
other merger stages when their morphological parameters are consistent with those of normal galaxies. The heterogeneous picture
of our ULIRGs thus may represent different evolutionary epochs of the ULIRGs.
In order to gain some insights into the $G-M_{20}$ morphology
we compare our $g$-band $G-M_{20}$ classification with the interaction
classification for the same sources made by Veilleux et al. (2002).
Veilleux et al. classified each object into one of the six sequential
merging stages according to its morphology. The comparison of two different
classifications is illustrated in Fig. \ref{morph.comp}. We find that all triple mergers by the
classification of Veilleux et al. are also classified as mergers by $G-M_{20}$.
Sources in their type IIIa group, i.e., wide binary pre-merger systems
with apparent separations greater than 10 kpc, are also very likely to be
recognized as $G-M_{20}$ mergers: $\sim 80\%$ of such sources are
recognized as mergers using $G-M_{20}$. Sources in their other groups
are less likely to be recognized by $G-M_{20}$ as mergers, and the fraction
of $G-M_{20}$ mergers appears to decrease toward later merging stages.
We find that $G-M_{20}$ mergers make up $\sim 31\%$ of the close binary pre mergers (type IIIb),
$\sim50\%$ of the compact mergers (type IVa), $\sim 46\%$ of the
diffuse mergers (type IVb), and only $\sim 11\%$ of the old mergers (type V)
are $G-M_{20}$ mergers, respectively. Sources classified as close binary
pre-merger systems (type IIIb) and old mergers (type V) have very large
fractions to be recognized as spiral galaxies by $G-M_{20}$ ($\sim 54\%$ and $\sim 67\%$,
respectively). These comparison results are qualitatively consistent with
the simulation results by Lotz et al. (2008) in which the most disturbed $G-M_{20}$ morphology
happens during the first pass and maximum separation. However, since our measured morphology is limited by
relatively small number statistics and relatively large uncertainties, we do not conclude
strongly that the observed morphology can be well explained by the simulations.
We also find that uncertainties in measuring the morphologies contribute significantly to the distribution of the ULIRGs in the parameter space.
LPM04 reported that the measurements of $G$ and $M_{20}$ are robust to within 10\%
when the source has signal-to-noise ratio per pixel $\langle S/N\rangle $ greater than 2.5,
the Petrosian radius larger than 2.5 times the PSF FWHM, and the physical resolution (parsec per pixel) higher than
500 $\rm pc\ pixel^{-1}$. However when the physical resolution is lower the uncertainties of morphologies largely increase
(Fig. 6 in their paper)
since small structures are washed out. At $z\sim 0.1-0.2$ the spatial resolution of the SDSS ULIRGs is roughly 0.7-1.3 $\rm kpc\ pixel^{-1}$
and the uncertainties need to be calibrated carefully.
We perform simple simulations to estimate the uncertainties of $G$ and $M_{20}$ for the SDSS ULIRGs at different noise levels.
We use four model elliptical galaxies, four spiral galaxies selected from SDSS images, and three merging galaxies
selected from our
ULIRG sample. The four elliptical galaxies are modeled using de Vacoulers profiles and are added to real SDSS images to mimic the observational
conditions. The three spiral galaxies are selected with bright $g$-band magnitudes of 13.9, 15.2, and 16.4, respectively, much brighter than the
median $g$-band magnitudes of the SDSS sample (18.0 mag). The three ULIRGs (FSC08572+3915, FSC14060+2919, FSC14121-0126)
are selected with bright $g$-band magnitudes and unmistakable merger morphology. The $g$-band magnitudes are 16.4, 16.7 and 17.8, respectively,
all brighter than the median g-band magnitudes of the ULIRGs (17.8 mag).
For each image we generate random noise at a series of different levels, and at each level we generate 20 noise maps independently
and add them to the original image.
Morphologies, Petrosian radius, and $\langle S/N\rangle$ are measured for each noise-added source and compared to the original measurements.
Although our simulation is limited by the sample size, $\sim90\%$ of the noise-added sources selected
by the signal-to-noise and size criteria ($\langle S/N\rangle>2.5$ and $R_p>2.6\arcsec$) have $g-$band magnitudes between 16 and 19,
and $g-$band Petrosian radii between 3 and 12 arcsec ($\sim$ 6 - 22 kpc at $z\sim0.1$), ranges similar to those of the ULIRGs.
We find that the measured Petrosian radius and the $\langle S/N\rangle$ do not always decrease monotonically with the amount of noise
added, and they are broadly distributed at low signal-to-noise levels. We also find that, rather than remaining fixed for a given source,
the measured $G$ decreases and $M_{20}$ increases systematically with increasing noise. Therefore when seen against higher
background noise, a galaxy tends to move to the lower left corner on the $G-M_{20}$ plot.
This effect will decrease the fraction of the mergers observed in the merger region.
The Gini coefficient $G$ decreases by as much as 0.2 with increasing background noise until the galaxy is no longer detected
(either $r_p < 2 \times FWHM$, or $\langle S/N\rangle < 2.5$), and $M_{20}$ increases
by as much as 0.5. The standard deviations of $G$ at every noise level are all less than 0.05,
and the spiral galaxies have the smallest dispersions. The standard deviations of $M_{20}$ are all less than 0.3-0.4,
and the elliptical galaxies have the smallest dispersions. There is no systematic difference for the uncertainties between different bands.
These uncertainties ($\Delta G=0.05$ and $\Delta M=0.5$) are shown in Fig. \ref{gm20} as error bars, and the systematic
trend with increasing noise in $G$ and $M_{20}$ is shown as an arrow.
The aperture within which the morphologies are measured is also an important source of
uncertainty. As noted by Lisker (2008), $G$ value measured within larger apertures have systematically larger
values, and the measured uncertainties are minimized when the Petrosian radius is used as the aperture size.
The reason behind is that more low surface brightness pixels are included within a larger aperture size,
which steepens the intensity distribution of the pixels and consequently increases the value of $G$.
Some of the low surface brightness features of the source, often at larger distances from the center, are mistakenly excluded
when the aperture size is chosen too small, which flattens the pixel intensity distribution and consequently decreases the value
of $G$. When assigning pixels to the sources we adopt the surface brightness at the Petrosian radius as the threshold, and
assign pixels brighter than the threshold to the source. Therefore, according to their study our measured morphologies should
have minimized uncertainties, because our apertures should be very close to the Petrosian radius and not significantly
larger or smaller.
\subsection{Relations among Morphology, Optical Color, and FIR Luminosity}
The morphology, color-magnitude relation, and the infrared luminosity of ULIRGs present
a rather complicated picture of these systems. The huge infrared luminosity in our ULIRGs suggests
that their dominant powering sources should be obscured by dust. However,
the majority of them are optically bright and blue. This further
implies that dust is not uniformly distributed
and that significant amount of the unobscured stellar light is seen directly.
Indeed, as shown in Fig. 1, there are obvious color gradients in the ULIRGs, with their tidal features at large
distances appearing blue and their central regions appearing red. Lotz et al. (2008b) suggested
that merging systems are most disturbed in $G-M_{20}$ space and are located in the merger region in the $G-M_{20}$ plot
during the first pass when the tidal features are prominent.
After the first pass they will move to the late-type and irregular region (see Fig. 5 in their paper).
Thus we would expect the integrated colors of those disturbed systems
selected by $G-M_{20}$ to be relatively bluer than the ones with normal $G-M_{20}$ relations,
due to the blue colors of the tidal features. Recent hydrodynamic simulations of disc-disc
mergers also suggested that the infrared luminosity of ULIRGs will peak during the final
merger when enough metals and dust have been accumulated (e.g. Jonsson et al. 2006),
so one would also expect sources with higher infrared luminosity are preferentially found in the late-type and irregular region.
We test these hypotheses by plotting the $g-$band $Gini$ vs $M_{20}$ for the ULIRGs in Fig. \ref{other.relation},
with different symbol sizes representing their FIR luminosities and different symbol colors representing
their $^{0.1}g-^{0.1}r$ colors. There are no strong relations between the optical colors
and $G$ or $M_{20}$, except that optically redder sources are slightly more
preferentially located in the non-merger regions with lower $G$ values.
This is further illustrated in the small panel in the same plot, where we plot the color distributions of sources
with $G\ge0.55$, the median $G$ of the whole sample, and of sources with $G<0.55$.
The sources with lower $G$ values tend to distribute toward redder optical colors, although the reddest source
has a higher-than-median $G$ value (0.62). All AGN ULIRGs have $G$ values greater than 0.5.
There are no obvious correlations between the FIR luminosity and $G$ or $M_{20}$.
These results are broadly consistent with the scenario that the blue optical colors originate from more disturbed systems,
and these systems are not during the phase when the FIR emission peaks. However, as mentioned earlier, the
uncertainties in measuring $G$ and $M_{20}$ could be substantial, and the measured morphologies are affected by orientation,
orbital parameters, viewing angle and dust content (Lotz et al. 2008b). We thus do not conclude any strong relations
among the morphologies, the FIR luminosity, and the optical color of these system.
\section{SUMMARY}
\label{summary}
We present color-magnitude and morphological analysis of 54 ultraluminous infrared
galaxies (ULIRGs) in the Sloan Digital Sky Survey (SDSS). The sample consists of all the ULIRGs from
the IRAS 1Jy sample (Kim \& Sanders, 1998) that have been imaged in the Data Release 5 of SDSS, spanning a
redshift range from 0.018 to 0.265 with a median redshift of 0.151.
The main results are:
$\bullet$ The ULIRGs are a very luminous group of galaxies in the optical, and the majority ($\sim93\%$) are more luminous than the
median r-band absolute magnitude of the SDSS comparison galaxies in the same redshift range.
$\bullet$ The ULIRG sample forms a distinct group in the color-magnitude diagram. 24 out of the 52 unsaturated ULIRGs ($\sim46\%$)
lie outside the 90\% level contour of the SDSS galaxies in the same redshift range (NYU-VAGC, DR4; Blanton et al. 2005).
The majority of the ULIRGs ($\sim87\%$)
have typical colors of the blue cloud, only $\sim6\%$ are located in the green valley, and $\sim7\%$ are located in the red sequence,
which implies that most of the ULIRGs are still undergoing copious star formation.
$\bullet$ There are 14 ULIRGs known to harbor an AGN (AGN ULIRGs). None of them are located in the green valley.
We use two simple approaches to estimate the point source flux in these systems, and find that on average the central
point source contributes less than one-third to the total luminosity in $r$ band. The host galaxies do not have a significantly
different distribution in the color-magnitude diagram after the point sources removed.
$\bullet$ We perform $G$ and $M_{20}$ analysis on the ULIRGs and find that their distribution in the $G-M_{20}$ space is heterogeneous.
Less than $\sim50\%$ (e.g. $\sim42\%$ in $g$ band) of the ULIRGs are located in the merger region found for local ULIRGs (LPM04). However the heterogeneous
distribution is consistent with the morphology produced by numerical simulations (Lotz et al. 2008b), and we also discuss the uncertainty
in the morphology measurements at the physical resolution of SDSS.
$\bullet$ We find that the ULIRGs have comparable optical luminosity as the SDSS QSOs within the same redshift
range but much redder. In comparison, the ULIRGs are much more luminous than the SDSS type 2 quasars within
the same redshift range and much bluer. The distribution of the SDSS type 2 quasars peaks at the green valley.
We perform two-dimensional K-S tests and the results show that the ULIRGs, the SDSS QSOs, and the SDSS type 2
quasars are statistically different samples in the
color-magnitude two-dimensional space.
$\bullet$ We study the relations among morphology, optical color, and far-infrared luminosity of the ULIRGs. These results
are broadly consistent with the scenario that the blue optical colors originate from more disturbed systems, and these systems
are not during the phase when the far-infrared emission peaks. However, the morphology parameters are
affected by the measurement uncertainties, as well as physical properties of the merger.
We thus do not conclude any strong relations among morphology, far-infrared luminosity, and optical color of these system.
Selected by their true infrared luminosity with no SED extrapolations, and with the uniformity of the SDSS optical photometry,
our ULIRG sample is a best low-redshift comparison sample to study the color-magnitude relation of high-redshift ULIRGs
and sub-millimeter galaxies. Although the morphological study of our sample is limited by the physical resolution of SDSS, the
size and structures of a galaxy might also be smaller at high redshift, and therefore our sample also provides a comparison
to study the morphology of high-redshift ULIRGs and sub-millimeter galaxies. We will discuss the color-magnitude relation
and morphology for a sample of the most luminous infrared galaxies at $z\sim1$ in a succeeding paper.
\acknowledgements
The authors would like to thank Dr. Jennifer Lotz for the use of the $G-M_{20}$ code.
Yuxi Chen would like to thank Dr. Daniel McIntosh for inspiring comments and suggestions. We also appreciate
some useful discussions with Prof. David Sanders and Prof. Nicholas Scoville.
This material is based upon work supported by the National Science Foundation under Grant No. NSF 0332504.
Funding for the creation and distribution of the SDSS Archive has been provided by the Alfred P. Sloan Foundation,
the Participating Institutions, the National Aeronautics and Space Administration, the National Science Foundation,
the U.S. Department of Energy, the Japanese Monbukagakusho, and the Max Planck Society. The SDSS Web site is http://www.sdss.org/.
The SDSS is managed by the Astrophysical Research Consortium (ARC) for the Participating Institutions.
The Participating Institutions are The University of Chicago, Fermilab, the Institute for Advanced Study, the Japan Participation Group,
The Johns Hopkins University, the Korean Scientist Group, Los Alamos National Laboratory, the Max-Planck-Institute for Astronomy (MPIA),
the Max-Planck-Institute for Astrophysics (MPA), New Mexico State University, University of Pittsburgh, University of Portsmouth,
Princeton University, the United States Naval Observatory, and the University of Washington.
|
\section{Motivation}
With the arrival of many-core architectures, the variance of
processors increases by another order of magnitude. This variance
increases also the need for high-level language virtual machines (VMs) to abstract
from variations introduced by differences among many-core architectures\cite{CellBEA,
Larrabee, GodsonT, Tile64}.
We are concerned with processors having multiple cores, using non-uniform memory access
architectures, and explicit mechanisms for inter-core communication.
For software developers, VMs have to provide abstractions from concrete hardware details
like number of cores or memory access characteristics. In the following subsection,
we categorize three groups of hardware architectures,
which need to be supported by VMs, as \emph{concrete concurrency models}.
In contrast to those concrete concurrency models, we refer to the concurrency models
defined by languages or libraries and used by application developers as
\emph{abstract concurrency models}.
Our claim is that the currently available incarnations of abstract concurrency models
in the form of languages and libraries are not sufficient and need to be complemented
by inherent support for multiple concurrency models by VMs.
To motivate our proposal, we analyze the challenges for VMs with regard to concrete as well as
abstract concurrency models in the remainder of this section.
The remainder of this paper discusses our idea of an instruction set for concurrency
and the research that has to be conducted to develop a methodology which allows
to tailor such an instruction set for the needs of a specific VM and its
application domain.
We give a brief overview of our
initial experiments and present the conclusions for a full-grown experimental
environment. We also discuss the related work which contributs approaches and solutions
to VMs for many-core architectures.
\subsection{Challenges for VMs on Modern Processor Architectures}
\label{sec:ChallendesModernProcessors}
Since processor vendors reached an upper bound on the possible clock speed
to gain more performance, the design of modern processor architectures
diverges from their predecessors in central design elements with each new generation,
trying to achieve better performance by introducing support for explicit concurrency.
This trend has much different consequences compared to the gradual architectural
changes over the last decade.
Instead of increasing the complexity of the memory hierarchy to hide latency
and bandwidth issues, introducing out-of-order execution of instructions, or
simply raising clock rates, changes are made which are not transparent to software anymore
and require special support. As detailed in the remainder of this section, the \emph{memory access
characteristics} change, the explicit concurrency increases the need for \emph{cache-conscious}
programming, and some architecture introduce explicit inter-core communication which all
needs to be support by VMs.
As already mentioned before, we refer to the concurrency models provided at the
hardware level as \emph{concrete concurrency models}. We identified three models and the challenges they imply for the implementation of VMs.
\subsubsection{Single-core Processor}
The most fundamental \emph{concrete concurrency models}
is a single-core system accessing memory not shared with
another processor. In such a system, the only notion of concurrency is provided
by the operating system (OS) offering some form of preemptive thread scheduling.
Modern single-core architectures usually use mechanisms
like out-of-order execution of instructions,
vector instructions, or pipeline steps which can lead to parallel execution of small
code portions. However for VM implementations, these forms of parallelism do not impose additional
complexity. It is not necessary to introduce a concurrent garbage
collector, but a just-in-time (JIT) compiler could still benefit from these mechanisms.
However, for optimal performance, these architectures put another burden
on programmers. Deep cache hierarchies have to be treated carefully for
optimal performance, i.\,e., programmers have to be \emph{cache-conscious}.
Thus, they are responsible for reorganizing data layouts to avoid phenomena like
cache thrashing and support the prefetching heuristics.
JIT compilers could actively use characteristics
like cache line sizes, prefetching heuristics, and branch prediction of the
various hardware architectures for optimization\cite{InstructionScheduling,
ArtOfMpP}, and interpreters could be adapted, e.\,g., to
assist hardware branch prediction\cite{SuperInstructions}.
With respect to concurrency provided by the OS, a VM has to define
a memory model\cite{1229469} and a task model.
The memory model specifies, amongst others, when a write to a shared variable by one thread can
be seen by reads done by another thread.
These guarantees interact in various ways with JIT compiler optimizations, like
storing temporary values in registers, and OS thread scheduling, since the guarantees
need to be enforced before a thread can be rescheduled. The best performance is
usually achieved if guarantees are less strong and provide opportunities for
reordering to hide memory latency.
The task model makes concurrency available
to language developers and should allow to schedule tasks with respect to the used data,
to use caches efficiently if tasks, e.\,g., in the form of threads, operate
on shared data.
\subsubsection{Multi-core Processor}
The second concrete concurrency model is a shared memory approach for multi-core
or hardware multi-threaded systems.
To allow a clear distinction to many-core processors (see below), we will concentrate on
systems with an architecture for uniform memory access (UMA)\footnote{Often UMA systems
are regarded as symmetric multiprocessing (SMP) systems, however, for this
discussion, the memory architecture is the main point of interest and the actually
utilizations of the cores is subordinated.},
i.\,e., multiple cores or threads connected to a single main memory system
and a cache hierarchy which provides cache coherency.
These architectures have grown from single-core processors and usually share all
important characteristics like deep cache hierarchies and out-of-order execution.
The main difference is the additionally provided hardware concurrency
and cache coherence.
The guarantees given by the memory model are even more important in this case.
Here it is not only arbitrary interleaving but parallel execution which has to
be taken into account.
Overly strict guarantees will require that writes are followed by memory barriers to
ensure that neither instruction-reordering nor the cache hierarchies are hiding
changes at any given time. This will of course hurt performance since both mechanisms
could be practically disabled.
By introducing cache coherency, the appropriate utilization of the available
hardware mechanisms becomes more complex. One example is given by Herlihy and
Shavit\cite{ArtOfMpP}. They discuss different lock implementations with the basic
insight that a synchronizing operation like \emph{compare-and-swap} provided by
the processor might hurt performance if used inappropriately. Combined with a simple
read operation which checks whether the value has changed utilizing caching,
performance can be improved, since relying on cache coherence has less overhead
than an operation which might need to synchronize different cores explicitly
and causes memory operations which cannot be cached.
This insight is not only important for the implementation of synchronization
primitives provided by VMs, but also for the implementation of JIT compilers to
generate efficient code.
Similar to single-core systems, task scheduling should respect data
dependencies.
For multi-core systems, scheduling should also be aware of
the cache architecture, i.\,e., how cores share caches and how caches are connected
to a hierarchy, to avoid cache thrashing or rather exploit caching efficiently.
\subsubsection{Many-core Processors}
In contrast to multi-core processors, many-core processors
cannot rely on a UMA architecture anymore since the known mechanisms do not scale\cite{1467875}.
Instead, these processors rely on non-uniform memory access
(NUMA) architectures, i.\,e., the cost to access a specific memory location
can be different for all cores.
Furthermore, some architectures will provide explicit communication
facilities between cores and thus will not rely solely on shared memory
for direct communication.
Others will try to avoid this additional complexity.
However, many-core architectures which provide shared memory and
coherent caches will exhibit performance behavior which will vary with respect to
data locality.
We will discuss three candidates from this category briefly.
\paragraph{Cell BE}
The Cell BE\cite{CellBEA} is already in wide use for media systems as well
as for scientific computing.
One of the major characteristics of the Cell BE is its heterogeneous approach
to combine a central processing element with multiple \emph{synergistic processing
elements} (SPE) to offload computational intensive tasks.
The SPEs are very simple and are not part of a cache hierarchy, do not feature
out-of-order execution, or even branch prediction. Each one has a local storage
but cannot access main memory directly. Instead, a SPE has to request blocks of
memory to be copied into its local store before it can use the data.
The interconnection of these cores is realized by a ring bus architecture.
Here the physical locality is important to achieve optimal performance.
The ring bus is build from four rings, where two rings can transfer data clockwise
and the other two can transfer data counter clockwise.
A more detailed overview of this architecture is given by Krolak\footnote{\url{http://www.ibm.com/developerworks/power/library/pa-fpfeib/}, Version: 29 Nov 2005}.
\paragraph{TILE64}
The processors produced by Tilera, e.\,g., the TILE64\cite{Tile64} are
somehow similar to the SPEs with regard to their simplistic design.
However, the TILE64 is a homogenous system with only one type of cores.
Each of the 64 cores has a small cache and is interconnected with neighboring cores
(tiles) via a mesh network with five independent special purpose networks.
Thus, to access memory, a core uses the \emph{memory dynamic network} which transports
the request to the according memory controller and returns the data.
Furthermore, an \emph{inter-cache network} allows to access the local caches of other
cores. Additional inter-core communication networks allow various direct communication
schemes between cores.
The challenges to implement VMs on top of such a system have been documented by
Ungar and Adams\cite{ManyCoreObjectHeap}. The crucial obstacles they encountered
where very small local caches, inefficient communication due to shared memory
(as opposed to explicit core-to-core communication), and
required replication of immutable objects to be cached locally since the processors
cache coherency protocol allows caching of a page only on its home core.
From these observations, we conclude that
adequate strategies will be required to implement object heaps enforced by very small caches, as well
as an appropriate way to harness the available bandwidth for inter-core communication
to reach the theoretical performance maximum.
\paragraph{Larrabee}
Intel's Larrabee\cite{Larrabee} represents another possible homogenous design.
Similar to the other two designs, the cores itself are much simpler than, for instance,
the latest designs used in desktop computers.
They use an in-order architecture extended by wide vector processing units
since it is primarily designed as a graphics processor.
However, in contrast to the other designs, Intel has decided to go with a cache coherent
system to hide some of the complexity. Each core has its own local subset of
the L2 cache and accesses main memory via the coherent L2 cache using a ring network.
At the moment, it seems that they will not expose this ring network explicitly and
communication is only done via shared memory.
Nonetheless, the performance characteristics will differ drastically from standard
multi-core system especially for systems with more than 16 cores where multiple
short linked rings will be used.
\subsection{Challenges for Abstract Concurrency Models}
\label{sec:MotivationAbstractModels}
Today's \emph{abstract concurrency models} are commonly regarded as not ideal
and a lot research is conducted to improve this situation with different approaches.
In short, shared memory with locking is too complicated \cite{ProblemWithThreads}
and software transactional memory (STM) \cite{STM} as well as Actors \cite{ActorsFormalism,
Actors} are promising but not widely adopted.
Thus, we expect that ongoing efforts in building languages,
to handle the inherent concurrency of many-core systems,
will likely lead to domain-specific languages and will require support by the underlying
VMs.
In this regard, VMs like the \emph{Java Virtual Machine} (JVM) and the
\emph{Common Language Infrastructure} (CLI) are
becoming more important as common execution platforms for multiple languages,
since not only the implementation of JIT compilers and efficient garbage
collectors is a tremendous effort, but
the ability to reuse the existing infrastructure surrounding a VM is an
economical concern.
Realistically, there will not be one single model for expressing concurrency.
Thus, we argue that a VM has to provide support for a wide range of concurrency models
at its core. Very likely, developers will have to deal with several models;
e.\,g., in relation with legacy code requiring proper support.
Furthermore, support for a wide range of models eases the work of language
designers to implement new ideas or domain-specific solutions. VM developers can also
benefit from richer concurrency semantics, as it would enable efficient
implementations of different
abstract concurrency models on top of the concrete models.
To illustrate our argument, we will discuss the example of the actor model \cite{ActorsFormalism}.
The JVM and CLI are both widely used and host all kinds of different programming
models. Functional as well as imperative languages and in the recent past they started
to provide support for dynamic languages, too.
However, if it comes to concurrency, both support only a shared-memory model with
threads and locks.
The implementation of an actor-based concurrency model, like it
is found for instance in Erlang\cite{Erlang}, on top of these VMs
has turned out to be a tough problem.
Karmani et al.\cite{1596658} surveyed different language and library implementations
of actor models on top of the JVM. They observed that only few of them actually
implement a model which preserves properties like isolation so that actors never share any
state in terms of references to a common object graph.
The few ones which do, usually rely on inefficient mechanisms like serializing the
object graph which is then send as a copy. A VM could provide support for
much more efficient zero-copying strategies and enforce the desired properties
of the actor model at the same time.
\subsection{Conclusions}
The presented concrete concurrency models represent actual hardware architectures
which differ widely. The important characteristics are their cache hierarchies,
memory access architectures, the provided form of concurrency,
and means for communication between cores.
Theses characteristics influence not only various implementation details all over
the VM but affect the optimal design of memory, task, and communication model for
each of the different concrete concurrency models.
For example, the challenge for VMs on many-core architectures is not solely the
utilization of available hardware concurrency but also to use the provided memory
and communication facilities appropriately.
Thus, VMs' concurrency abstraction layers must enable efficient implementations on
top of the different concrete concurrency models.
To achieve that,
we argue that VMs should provide explicit and comprehensive support for concurrency.
Explicit support for the various different abstract concurrency models
would allow direct mappings from congruent models which will allow an efficient
utilization of the available facilities and would ease the task to find a suitable
mapping for the remaining, not directly supported concepts.
For instance, the discussed actor model offers opportunities
for an efficient mapping onto many-core architectures.
Since cache coherence is an issues in these architectures, it would
be possible to use shared-memory only for immutable global state.
The state of single actors could be stored in distinct parts of the memory,
so that false sharing is avoided and the small local caches can reach
peak efficiency.
In a standard JVM, it would be rather hard to reconstruct the necessary semantics
for such a mapping from the bytecode, but a semantically enriched instruction set
could would allow a JIT compiler to apply such optimizations.
\section{VM Instruction Sets with Concurrency Support}
Our proposal to achieve a concurrency abstraction layer is to extend the VM instruction
set by concurrency operations. Such an
instruction set will decouple the concurrency models on the different levels of implementation
in such a way that they can be varied independently.
Fig.\,\ref{abstraction-layer} visualizes this idea by showing three different abstract
concurrency models mapped to an instruction set with explicit concurrency support implemented
on top of three different concrete concurrency models.
\begin{figure}[hbtc]
\begin{center}
\includegraphics[width=0.8\textwidth]{figure1.pdf}
\caption{A VM instruction set as abstraction layer between abstract and concrete concurrency.}
\label{abstraction-layer}
\end{center}
\end{figure}
Expressing concurrency in the instruction set instead of using
libraries has two major advantages. First,
it will be possible to compile concurrency-related language constructs directly to these
instructions, avoiding dependencies between languages and libraries on top of the VM.
Second, this choice leads to a larger optimization potential at the VM level, e.\,g., for
JIT compilation, which benefits from the instruction set's precise semantics.
Since there will not be a single instruction set matching all possible requirements, we will work on one
instruction set representing a very generic
set of requirements, and investigate the design tradeoffs to derive design advice for more concrete
requirements as well. Thus, we plan to devise
a methodology to develop VM instruction sets with inherent concurrency support, enabling
VM designers to build a concurrency abstraction layer optimized for their
particular requirements.
The methodology will describe how to decouple
abstract and concrete concurrency. Language designers will be provided with a strategy to
map abstract concurrency models to instruction sets, and VM implementers
will be enabled to implement instruction sets efficiently on top of concrete
concurrency models.
The methodology will not only guide such undertakings, but will also give an impression on the effort necessary for their realization.
Below, we discuss our approach in more detail.
\subsection{Approach to Synthesize the Instruction Set}
To devise a broadly applicable methodology, we decided
to adopt a step-wise approach to designing a general instruction set and
discovering the important design tradeoffs. The three currently
most important concurrency models are significantly different in how they represent and realize concurrency: shared-memory with locking, STM, and actors.
For each of these, we will survey different
incarnations in languages or libraries to find each model's
set of primitives relevant for an instruction set.
Potential candidates for examination are, to name just a few, Java\cite{JavaSpec},
C\#\cite{Ecma334}, Smalltalk\cite{165375}, Cilk/JCilk\cite{Cilk, JCilk}, and frameworks like
Fork/Join for Java\cite{JavaForkJoin}.
Furthermore, constructs like monitors and semaphores are considered as well\cite{506118}.
In the field of STM, we currently consider the work of Shavit and Touitou\cite{STM},
Ziarek et al.\cite{UniformTM}, Saha et al.\cite{1123001}, and Marathe et al. \cite{1066660}.
In the world of actor models the work of Hewitt et al.\cite{ActorsFormalism} and Agha\cite{Actors}
as well as the languages Erlang\cite{Erlang}, Scala\cite{UnifyThreadsEvents},
and Kilim\cite{Kilim} are considered starting points.
\subsection{Ideas to Combine Abstract Concurrency Models}
One of the major research challenges will be to find appropriate combinations of
the different abstract concurrency models. The idea is not to build an
instruction set which is a simple enumeration of primitives for the different
models, but instead an elaborated combination thereof. Thus, the interaction
between different models has to be completely understood and defined, too.
Our ideas for model combinations are based on the following work.
Volos et al.\,\cite{LocksAndTM} and Blundell et al.\,\cite{UnrestrictedTM} have
described possible solutions for combining locking based code with STM.
A combination of locking based code and actors is described by
Van Cutsem et al.\,\cite{1352693}. STM has many similarities with common transaction processing systems; thus, we will
investigate the application of transaction processing monitors
\cite{TransactionProcessing} as used in distributed settings to use STM in conjunction with actors.
\subsection{Tradeoffs to be Investigated}
For the methodology, the discussion of the following design tradeoffs will be an important
part.
\begin{description}
\item[Model Combination:] Different solutions to combine concurrency models on the instruction
set level will be considered,
and their benefits and drawbacks investigated. This will reveal
critical details like incompatibilities and the possible degree of concurrency.
\item[Model Mapping:] Strategies to map the concurrency models preserved
in the instruction set onto concrete concurrency models.
Here the differences in the memory models, cache hierarchies, and communication
mechanisms have to be considered.
\item[Condensed vs. Bloated Instruction Set:] Only few instructions should be added
to avoid exceeding the limited number
of instructions in a typical bytecode set. However, additional semantics
in the instruction set could reduce the complexity of implementing
an abstract concurrency model on top of it. It can also be beneficial
for an efficient mapping to a concrete concurrency model.
Since language and VM implementations should be reasonably manageable,
these conflicting interests have to be
investigated.
\item[Bytecode vs. High-level Representation:] Currently,
bytecode sets are the most common representation for VM instruction
sets. With respect to communication centric many-core architectures,
we will investigate the potential of abstract syntax tree-like high-level
representations of interpretable code
in terms of reducing the implementation effort for new instructions
and JIT compilers. These investigations will be based on the work of
Kistler and Franz\,\cite{TreeJava}.
\item[Instruction Set vs. Standard Library:] A strategy, to decide which concepts
are valuable in the instruction set itself, e.\,g.,
by facilitating JIT compiler optimizations, and which are less common
or less fundamental and should be provided only in the standard library
for a given application domain, is necessary, too.
\end{description}
\section{Initial Experiments}
For our first basic experiments we used
SOM++\footnote{\url{http://hpi.uni-potsdam.de/swa/projects/som/}},
a very simple VM, implementing a Smalltalk-like language.
This VM is designed to be used for teaching and to prototype ideas rapidly.
Originally, it has a very small instruction set (16 instructions) and features a
straightforward bytecode interpreter. Its overall design favors simplicity over performance
and utilizes C++ to provide an object-oriented implementation. This results in a VM
implementation which emphasizes conceptual clarity. Thus, experiments usually require
a minimal effort. The downside of this approach is, that SOM++ is considered unoptimized with
regard to performance. Hence, experiments on SOM++ are useful to show the general impact of
different implementation strategies, for instance for garbage collection, but only provide a
rough estimate about performance and interaction effects between subsystems.
In the context of our first experiments, this is not an issue. The goal was to gain an impression
of the general impact of introducing concurrency related instructions into the bytecode
set of a virtual machine. In our experiments, we chose to focus on shared memory and non-shared
memory concurrency in the first place.
The foundation for these experiments is the SOM++ bytecode set. As mentioned before, it consists of 16
instructions. It is purely stack-based and design with simplicity as the main goal in mind.
Thus, the bytecodes are encoded as bytes with the values from 0 to 15.
Even though it would be possible to encode arguments---e.\,g., indexes for local variables or symbols---within the remaining bits, they are provided as an additional byte each.
Thus, bytecode instruction length varies in the range from 1 to 3. The bytecode set is outlined in Tab.\,\ref{SomBytecodes}.
\begin{table}[htdp]
\begin{center}
\begin{tabularx}{0.95\linewidth}{lX}
\lstinline|DUP| & duplicate top element \\
\lstinline|PUSH_*| & push locals, arguments, fields, blocks, constants, and globals onto stack \\
\lstinline|POP| & remove top element \\
\lstinline|POP_*| & pop top element to locals, arguments, and field variables\\
\lstinline|SEND sig| & send a message identified by \lstinline|sig| to the top element \\
\lstinline|SUPER_SEND| & send a message to the top element, use implementation of the parent class\\
\lstinline|RETURN_LOCAL| & return from the current block of execution to its outer context\\
\lstinline|RETURN_NON_LOCAL| & leave the currently executed method from an inner block\\
\lstinline|HALT| & leave the interpreter loop
\end{tabularx}
\caption{SOM++ bytecode set}
\label{SomBytecodes}
\end{center}
\end{table}
In the following sections, we briefly describe the two experiments, to illustrate potential concurrency
related instructions in VM bytecode sets.
\subsection{Shared Memory Concurrency}
Our very first experiment was to add basic instructions for shared-memory concurrency to
the SOM++ bytecode set.
We designed the extension similar to the existing instructions. Simplicity was not the foremost
concern here, but we have chosen to add only the five basic instructions outlined in
Tab.\,\ref{SomThreads}.
They operate on the top element of the execution
stack. For \lstinline|SPAWN|, the top element has to be a block which is then executed in a new thread. As a result, \lstinline|SPAWN| pushes a new thread
object onto the stack. The other four operate on an arbitrary object on the top of
the stack. The stack itself is not affected.
We relied on an existing implementation of
shared-memory concurrency using the \emph{Pthreads} library. Thus, the largest part of
the work was refactoring the existing implementation from primitives, i.\,e.,
native functions for the Smalltalk thread library to bytecode instructions.
Subsequently, the SOM++ compiler was adapted to emit the new bytecodes on
special messages.
\begin{table}[htdp]
\begin{center}
\begin{tabularx}{0.69\linewidth}{lX}
\lstinline|SPAWN| & spawn a new thread with the given block on top of the stack \\
\lstinline|LOCK| & lock the lock of the top element \\
\lstinline|UNLOCK| & unlock the lock of the top element \\
\lstinline|WAIT| & wait on a notification on the top element \\
\lstinline|NOTIFY| & notify all threads waiting on the top element
\end{tabularx}
\caption{Additional instructions for shared memory concurrency}
\label{SomThreads}
\end{center}
\end{table}
In the context of SOM++, the question arose whether it is beneficial to have these instructions in
the instruction set instead of implementing them as primitives. In the course of this project, bytecode instructions are actually the only option, which however brought about considerable overhead in implementing the required extensions in the compiler.
However, SOM++ is not the type of virtual machine we like to target with such
extensions. Instead,
these kinds of instructions are meant for multi-language virtual machines. Here, the
purpose of
an instruction set shifts from being a runtime representation of a program to being a
full-fledged assembly language for all kinds of language implementations.
Thus, a richer instruction set allows to move implementation effort from the
language-level, which has
to be redone for each language, to the platform-level where all language
implementations can benefit from it without additional effort.
For future experiments we will consider additional shared-memory operations to
increase the flexibility and expressiveness of the instruction set.
At the moment, we think that several low-level operations known from hardware
instruction set architectures could be useful additions to allow language designers
for instance to use lock-free synchronization mechanisms or data structures
at the heard of their languages.
Examples for such operations are atomic updates like \lstinline|XADD|
and \emph{compare-and-swap} (\lstinline|CMPXCHG|) from the IA-32 instruction
set architecture\cite{PentiumSpec}, as well as operations like
\emph{load-and-reserve}/\emph{store-conditional} which are included in the
PowerPC instruction set architecture\cite{PowerPCSpec} in form of \lstinline|lwarx|
and \lstinline|stwcx|.
\subsection{Non-shared Memory Concurrency}
The second experiment we conducted was inspired by the work of Schippers et\,al.\,\cite{TowardsACMM}
describing an actor-based machine model.
The aim of this experiment was to adapt SOM++ to implement concurrency by actors which do not share
memory, but use explicit message passing for communication.
This kind of machine model is typically found in distributed object systems\,\cite{232711}.
In this model, actors are containers for objects. It is derived from the notion of
\emph{vats} introduced in the E language (and its predecessors)\cite{ELangActors}
where actors are not ``active objects'', but containers for a number of
regular objects.
The contained objects are shielded from undesired concurrent modifications,
since each actor only has a single thread of control.
Messages between actors are exchanged using an incoming message queue per actor.
Objects can reference objects located in another actor by means of remote references.
Usual message sends between objects can be synchronous or asynchronous, independent from
whether the message is sent locally or over a remote reference.
Inside an actor, coroutines are allowed to support a simple means of concurrency.
This is useful since synchronous message sends over remote references
do not block the sending actor, but can yield control to another coroutine until
the return message is received.
To support this machine model, the instruction set had to be adapted as outlined in
Tab.\,\ref{SomActors}. The basic instructions
stay the same except for \lstinline|SEND|.
For message sends to objects over remote pointers, \lstinline|SEND| was adapted.
It forwards the message sent to the actor owning the object and yields the coroutine to wait
for the result value. The result value is later returned by the \lstinline|RETURN_REMOTE| bytecode.
Usual asynchronous message sends are realized by the \lstinline|SEND_ASYNC| bytecode and
coroutines can explicitly yield control using the \lstinline|YIELD| bytecode.
\begin{table}[htdp]
\begin{center}
\begin{tabularx}{0.9\linewidth}{lX}
\lstinline|SEND| & sends of remote references yield coroutine and wait for return value \\
\lstinline|RETURN_REMOTE| & sends the return value to the waiting coroutine \\
\lstinline|SEND_ASYNC| & send a message asynchronously to an object, the message
queue of the actor owning the receiving object is used \\
\lstinline|YIELD| & yields control flow, possibly to another coroutine
\end{tabularx}
\caption{Additional instructions for non-shared memory concurrency}
\label{SomActors}
\end{center}
\end{table}
\subsection{Choosing a Research Platform}
From our experiments, we conclude four requirements
for a full-grown experimental environment fit to
demonstrate the advantages of an instruction set supporting a wide range
of concurrency models:
\begin{itemize}
\item The VM has to be portable to platforms like TILE64\cite{Tile64}
or Cell BE\cite{CellBEA} to be able to evaluate the benefits in
mapping from an extended instruction set to different concrete
concurrency models.
\item Implementations of considered abstract concurrency models
which use a compilation to the VM instruction set as
implementation strategy should be available.
\item The VM instruction set should provide space (i.\,e., unused bytecode instructions)
for experiments.
\item The VM should provide an easy to adapt
JIT compiler to experiment with optimizations.
\end{itemize}
Based on these requirements, we compiled a list of more than fifty VMs comparing
mainly open source
implementations for various languages like Erlang, JavaScript, Python, and Scheme.
Here we present only a small subset of this comparison to discuss the reasoning
for choosing our research platform.
Tab.\,\ref{VMOverview} lists for each VM the characteristics of interest to choose our
research platform.
The column \emph{language} contains the target language implemented by the VM,
\emph{ACM} reflects the abstract concurrency model.
The availability of threads, STM, and actors implementations are represented by \emph{IS} for
instruction set support, \emph{Lib} for available libraries or language implementations,
or \emph{``-''} if implementations are not available but described in literature.
Furthermore, we consider whether a JIT compiler is available and a port to a many-core system
would be feasible.
PyPy's thread support is marked with a ``-'', since it relies on a global interpreter lock and
thus does not allow true parallelism. This DisVM was included even so we only have
access to its specification.
\begin{center}
\begin{tabularx}{1.0\linewidth}{llccccccrl}
& & & & & & & & Size & Impl. \\
Name & Language & ACM & Threads & STM & Actors & JIT & Port & (SLoC) & Lang. \\ \hline
SOM++ & Smalltalk& T/L\footnote{T/L: threads and locks}
& IS & & IS & & x & 6k & C++ \\
Lua & Lua & & Lib & & Lib & & x & 13k & C \\
LuaJIT\footnote{\url{http://luajit.org/}}
& Lua & & Lib & & Lib & x & x & 20k & C \\
RVM\footnote{A Squeak VM developed at IBM Research\cite{ManyCoreObjectHeap} for the TILE64}
& Smalltalk& T/L & Lib & - & - & & x & 28k & C++\\
CacaoVM\footnote{\url{http://www.cacaovm.org/}}
& Java & T/L & Lib & Lib & Lib & x & x & 121k & C++ \\
Mozart\footnote{\url{http://www.mozart-oz.org}}
& Oz & Data-flow & & & & & & 159k & C++ \\
Erlang& Erlang & Actors & & & IS & x & x & 247k & C \\
PyPy & Python & T/L & - & Lib & Lib & x & & 318k & RPython \\
Maxine\footnote{\url{http://research.sun.com/projects/maxine/}}
& Java & T/L & IS & Lib & Lib & x & & 361k & Java \\
HotSpot& Java & T/L & IS & Lib & Lib & x & & 540k & C++ \\
JikesRVM& Java & T/L & IS & Lib & Lib & x & & 978k & Java \\
DisVM\footnote{\url{http://doc.cat-v.org/inferno/4th_edition/dis_VM_specification}}
& Limbo & CSP\footnote{CSP: Communicating sequential processes\cite{CSP}}
& & & & & & spec. &
\end{tabularx}
\end{center}
\captionof{table}{Overview of potential research platforms}
\label{VMOverview}
~\\
Starting with SOM++, we have to conclude from our experience, that its idealized architecture
and its simple implementation allows for fast prototyping of ideas, but on the other hand
might conceal problems associated with our approach especially with regard to performance.
Lua is also small, but has been implemented with a clearer performance objective.
Furthermore, an implementation with a JIT compiler exists
which is small enough to
be ported to a many-core architecture without requiring overly large effort.
Thus, we will consider it as a vehicle to validate our research in the context of
embedded VMs.
The RVM is already tailored to the TILE64 processor.
Since it utilizes the many-core architecture, its special inter-core communication
facilities, and has a moderate complexity, we will use it for our first experiments,
applying our idea in the setting of many-core systems.
CacaoVM seems to be the smallest and most widely ported open source JVM with a JIT compiler.
Compared to other JVMs in the table, a port of the CacaoVM to a many-core system
should be more feasible, especially since it already has been ported to the
Cell BE\,\cite{CacaoCell}.
However, it might become necessary to consider VMs like HotSpot, JikesRVM and Maxine when it comes
to the validation of performance properties. At the moment, it is still not clear whether
we will need a JIT compiler with production-level performance to rule out performance
characteristics not introduced by our approach but other modifications done in the development.
Erlang, Mozart, and the DisVM have been included for consideration since they
implement other abstract concurrency models than the usual shared-memory model with
threads and locks. Interpreted Erlang got already official support for the TILE64
and will allow to conduct partial experiments. However, due to its
nature of a VM for a functional language and the complexity of its JIT compiler, we will not
chose it as our main research platform.
Mozart implements an abstract concurrency model based on data-flow variables.
Due to its complexity and focus on distributed environments it does not seem to be a feasible
platform for our research.
The DisVM is an interesting design of a VM where the abstract concurrency model is inspired
by CSP. Unfortunately we do not have access to the implementation and thus, an evaluation
as a research platform was not possible.
\section{Related Work}
Support for concurrency in VM instruction sets is currently limited.
The Erlang VM's BEAM instruction set\footnote{\url{http://erlangdotnet.net/2007/09/inside-beam-erlang-virtual-machine.html}}
is a notable exception, providing dedicated support for its efficient light-weight process
implementation. It includes instructions for asynchronous message sends, reading from the
process' mailbox, waiting and timeouts. It is an example of how one particular model can
be supported at the core of the VM.
Another example is the DisVM. It provides instructions to create channels between non-shared
memory threads as well as to receive and send messages synchronously.
Still, we argue that this concurrency support is not sufficient,
since each VM only provides support for a single abstract concurrency model.
By contrast, today's VMs have to support many different programming models
to justify the investments in sophisticated and efficient JIT compilers and garbage
collectors. Thus, they have to provide the basic means for a wide range of concurrency
models in the same way as they act as execution platforms for different languages.
In the broader field of instruction set design, there are ongoing efforts to extend the
capability of the JVM to act as a platform for different programming languages by introducing
the \texttt{INVOKEDYNAMIC} instruction\footnote{\url{http://jcp.org/en/jsr/detail?id=292}}. More general work on improving instruction sets with semantic extensions\cite{997182, Peymandoust2003} has been done for the hardware level, but the
concepts for, e.\,g., compiler adaption can be applied to VMs as well.
For the Cell BE, VM applicability has been evaluated.
Besides porting and designing JVMs for this platform \cite{CellVM, CacaoCell},
some optimizations have been considered to utilize available computation power
\cite{1346276, 1366265}.
Distributing a VM over several computational elements
bears additional challenges. Some of them have been addressed for VMs distributed
on cluster setups; e.\,g., class loading, strategies for
distributed method invocation, data access on the VM level \cite{dJVM}, or thread migration
\cite{Jessica}.
\section{Summary and Future Work}
We proposed to decouple abstract and concrete concurrency models to be able
to cope with the variability of upcoming many-core architectures and their different
memory access architectures.
We argue that this step is necessary to be able to provide support for several kinds of
languages and their abstract concurrency models on top of a VM.
Furthermore, the benefits of a semantically rich concurrency abstraction layer will allow
more efficient VM implementations on the various different hardware platforms.
The goal of our ongoing research is to design a comprehensive methodology to design VM instruction
sets combining several concurrency models to provide this abstraction.
The methodology will address the various different design tradeoffs.
Our preliminary prototype enabled us to refine our initial requirements for an experimental
environment and provided us with the necessary insights to be able to proceed with our
research on a suitable platform.
The next step of our work is to investigate the design principles for intermediate languages
and the state of the art in concurrency support. Preliminary results on this work have
been presented at the workshop on \emph{Virtual Machines and Intermediate Languages} 2009\cite{VMIL09}.
With the insights of design tradeoffs for the languages, i.\,e., the instruction sets themselves,
we plan to investigate which low-level primitives for shared memory concurrency should be
included. Later, the integration with non-shared memory models in the same language will
be tackled and thus, we will do the step to real multi-model concurrency support for VMs.
\bibliographystyle{eptcs}
|
\section{Introduction}
\label{Introduction}
The most natural way of defining the entropy of a quantum system
is to use the von Neumann entropy:
\begin{equation}
S_{\rm vN} = - {\rm Tr}[\hat \rho\ln(\hat \rho)]\,,
\label{entropy:vN}
\end{equation}
where $\hat\rho$ denotes the density operator which in the
Schr\"odinger picture satisfies the von Neumann equation:
\begin{equation}
\imath \hbar\frac{\partial}{\partial t}\hat \rho = [\hat H,\hat
\rho] \,, \label{density op:eom}
\end{equation}
where $\hat H$ is the Hamiltonian. Since quantum mechanics and
quantum field theory are unitary theories, the von Neumann entropy
is conserved, albeit in general not zero.
The decoherence program~\cite{Zeh:1970,Joos:1984uk} usually
assumes the existence of some environment that is barely
observable to some observer. This allows us to construct a reduced
density operator $\hat\rho_{\mathrm{red}}$ characterising the
system S, which is obtained by tracing over the unobservable
environmental degrees of freedom E: $\hat\rho_{\mathrm{red}} ={\rm
Tr}_E[\hat \rho]$. This is a non-unitary process which
consequently generates entropy. This operation is justified by the
observer's inability to see the environmental degrees of freedom
(or better: access the information stored in the ES-correlators).
Yet, tracing out some degrees of freedom results in complications.
The simple looking von Neumann equation~(\ref{density op:eom}) is
replaced by an equation for the reduced density operator. This
``master equation''~\cite{PazZurek:2001,Zurek:2003zz} can be
solved for only in very simple situations, which has hampered
progress in decoherence studies in the context of interacting
quantum field theories. In particular, we are not aware of any
known solution of the master equation that would include
perturbative corrections and implement the program of
renormalisation. Also, one has no control of the error in a
calculation in the perturbative sense induced by neglecting
non-Gaussian corrections.
Here we propose a decoherence program that can be implemented in a
field theoretical setting and that also allows us to incorporate
renormalisation procedures. The idea is very simple, and we
present it for a real scalar field $\phi(x)$. A generalisation to
other types of fields, e.g. gauge and fermionic fields, should be
quite straightforward.
The density matrix $\hat{\rho}(t)$ contains all information about
the (possibly mixed) state a quantum system is in. From the
density matrix one can, in principle, calculate various
correlators. For example, the three Gaussian correlators are given
by:
\begin{subequations}
\label{3 equal time correlators}
\begin{eqnarray}
\langle \hat{\phi}(\vec{x}) \hat{\phi}(\vec{y}) \rangle &=& {\rm
Tr}[\hat \rho(t) \hat\phi(\vec x)\hat\phi(\vec y)] =
F(\vec{x},t;\vec{y},t')|_{t=t'}\label{3 equal time correlatorsa}
\\
\langle \hat{\pi}(\vec{x}) \hat{\pi}(\vec{y}) \rangle &=& {\rm
Tr}[\hat \rho(t) \hat\pi(\vec x)\hat\pi(\vec y)] =
\partial_{t} \partial_{t'} F(\vec{x},t;\vec{y},t')|_{t=t'} \label{3 equal time
correlatorsb}
\\
\frac{1}{2} \langle \{ \hat{\phi}(\vec{x}), \hat{\pi}(\vec{y}) \}
\rangle &=& \frac12 {\rm Tr}[\hat \rho(t)\{ \hat\phi(\vec
x),\hat\pi(\vec y)\}] = \partial_{t'}
F(\vec{x},t;\vec{y},t')|_{t=t'} \label{3 equal time
correlatorsc}\,,
\end{eqnarray}
\end{subequations}
where $\hat \pi$ denotes the momentum field conjugate to
$\hat\phi$, and where the curly brackets denote the
anti-commutator as usual. Note that these three Gaussian
correlators can all be determined from the statistical propagator
$F(\vec x,t;\vec y,t')$ in the Heisenberg picture as usual in
quantum field theory. The fourth Gaussian correlator is trivial as
it is restricted by the canonical commutation relations. There is
an infinite number of higher order, non-Gaussian correlators that
characterise a system, where one can think of e.g.:
\begin{subequations}
\label{NGCorrelators}
\begin{eqnarray}
\langle \hat{\phi}(\vec{x}_{1}) \cdots \hat{\phi}(\vec{x}_{n})
\rangle &=& {\rm Tr}[\hat \rho(t) \hat\phi(\vec{x}_{1}) \cdots
\hat\phi(\vec{x}_{n})] \label{NGCorrelatorsa}
\\
\langle \hat{\pi}(\vec{x}_{1}) \cdots \hat{\pi}(\vec{x}_{n})
\rangle &=& {\rm Tr}[\hat \rho(t) \hat\pi(\vec{x}_{1}) \cdots
\hat\pi(\vec{x}_{n})] \label{NGCorrelatorsb} \,,
\end{eqnarray}
where $n \geq 3$. Moreover, one can also imagine a higher order
correlation function consisting of a mixture of both
$\hat{\phi}$'s and $\hat{\pi}$'s. Finally, one could think of
other non-Gaussian correlators obtained by further differentiating
the statistical propagator:
\begin{equation}
\label{NGCorrelatorsc}
\partial_{t}^{n}\partial_{t'}^{m} F(\vec{x},t;\vec{y},t')|_{t=t'}
= \partial_{t}^n \partial_{t'}^{m} {\rm Tr}[\hat \rho(t_{0})
\hat\phi(\vec x,t)\hat\phi(\vec y,t')] |_{t=t'}\,,
\end{equation}
\end{subequations}
where $n+m \geq 3$. For free theories, all the non-Gaussian
correlators either vanish or can be expressed in terms of the
Gaussian correlators.
Let us turn our attention to interacting field theories and
present two simple arguments why, generally speaking, the Gaussian
correlators dominate over the non-Gaussian correlators. In
interacting quantum field theories, the Gaussian correlators all
stem from tree-level physics, whereas non-Gaussian correlators are
generated by the interactions. Relying on a perturbative treatment
of these interactions, we can safely expect that all of these
higher order, non-Gaussian correlators above are suppressed as
they generically are proportional to a non-zero power of the
(perturbatively small) interaction coefficient. A second, more
heuristic, argument invokes the (classical) central limit theorem,
according to which systems with many stochastic mutually
independent degrees of freedom tend to become normally
distributed, and hence nearly Gaussian. While this theorem is
derived for classical systems and for independent stochastic
variables, we believe that it will in a certain quantum disguise
also apply to weakly interacting quantum systems with many degrees
of freedom.
Hence, we expect that to a good approximation, many of the
relevant properties of quantum systems are encoded in the Gaussian
part $\hat \rho_{\rm{g}}$ of the density operator $\hat \rho$ and
the non-Gaussian correlators are thus typically much more
difficult to access\footnote{As a simple example, consider the
temperature correlations induced by scalar field perturbations
from inflation: while the amplitude of scalar field fluctuations
is given by ${\rm Tr}[\hat \rho(t) \hat\phi_N(\vec x,t)^2]\sim
G_NH^2\sim 10^{-12}$, where $\phi_N$ denotes the Newtonian
potential, where $G_N$ is Newton's constant and $H$ is the Hubble
rate during inflation, its quantum corrections are expected to be
of the order $\sim (G_NH^2)^2\sim 10^{-24}$ or smaller.}. The
crucial point is that all information about a general Gaussian
density matrix and thus about the Gaussian part of a state, is
stored only in the three equal time correlators (\ref{3 equal time
correlators}). All non-Gaussian correlators like
(\ref{NGCorrelators}) contain information about the correlations
between the system and environment as well as the environment's
correlations induced by higher order interactions. When projected
on the field amplitude basis in the Schr\"odinger picture, the
Gaussian part of a density matrix reads:
\begin{equation}
\rho_{\mathrm{g}}[\phi,\phi';t] \equiv \langle \phi| \hat
\rho_{\mathrm{g}}(t) |\phi'\rangle = {\cal N}\exp \! \left[\! - \!
\int \! \mathrm{d}^3x \mathrm{d}^3y\left( \phi(\vec x\,) A(\vec
x,\vec y;t)\phi(\vec y\,) +\phi'(\vec x\,) B(\vec x,\vec
y;t)\phi'(\vec y\,)-2\phi(\vec x\,) C(\vec x,\vec y;t)\phi'(\vec
y\,)\right) \right]\! , \label{density matrix: Gaussian}
\end{equation}
where $A,B,C$ and ${\cal N}$ are functions of the equal time field
correlators (for details see sections~\ref{Entropy in scalar field
theory}). Of course, the von Neumann equation~(\ref{density
op:eom}) does not hold any more for $\hat \rho_{\rm g}$, and thus
the entropy conservation law is violated by $\hat \rho_{\rm g}$.
Indeed, neglecting the information stored in observationally
inaccessible correlators will give rise to an increase of the
entropy\footnote{In passing we note that we shall not address
entropy generated due to the loss of unitarity caused by some
observer's inability to access the information stored in all of
the space-time. Prominent examples of this research field include
black hole and de Sitter space entropy and there is an extensive
literature on the subject. In \cite{Campo:2008ju} truncation of
the infinite hierarchy of Green's functions was proposed as an
operational definition of the entropy of cosmological
perturbations. Calzetta and Hu \cite{Calzetta:2003dk} coined the
expression ``correlation entropy'' and proved an H-theorem for the
correlation entropy in a quantum mechanical model.}.
The purpose of the present work is to calculate the entropy of a
system taking these considerations into account. We distinguish
two cases. In the simplest case we assume that our observer can
only measure Gaussian correlators and his apparatus is insensitive
to all non-Gaussian correlators. We will consider this situation
in sections \ref{Gaussian entropy from the Wigner function} and
\ref{Gaussian entropy from the replica trick}. In quantum field
theories, the problem seems to be that the evolution equations of
the three Gaussian correlators~(\ref{3 equal time correlators}) do
not close as soon as perturbative corrections are included. The
reason is that off-shell physics becomes important. A way out of
this impasse, and in our opinion {\it the} way out, is to solve
for the dynamics of the unequal time statistical correlator, i.e.
the statistical propagator, from which one then extracts the three
relevant correlators~(\ref{3 equal time correlators}).
In the second case, we assume that our observer is sensitive to
specific types of non-Gaussianities present in our theory, apart
from the leading order Gaussian correlators of course, either
because the observer has developed a sensitive measurement device
or the non-Gaussianities happen to be large enough. We will
consider this case in section \ref{Non-Gaussian entropy: Two
examples}. The sections above contain, for pedagogical reasons,
quantum mechanical derivations rather than quantum field
theoretical calculations. We will generalise our results to
quantum field theory in section \ref{Entropy in scalar field
theory}.
This paper aims only at developing the necessary machinery to
discuss entropy generation in an interacting quantum field theory
in an out-of-equilibrium setting, see e.g. \cite{Chou:1984es,
Jordan:1986ug, Calzetta:1986cq, Aarts:2001qa, Aarts:2001yn,
Aarts:2002dj, Berges:2002cz, Juchem:2003bi, Juchem:2004cs,
Berges:2004yj}. We discuss the application of these ideas
elsewhere \cite{Koksma:2009wa, Koksma:2010} in great detail, where
we consider a scalar field $\phi$ coupled to a second scalar field
via the interaction Lagrangian density ${\cal L}_{\rm int} =
-\frac{1}{2} h \chi^2(x) \phi(x)$. The results we develop here are
applicable to a much more wider class of field theoretical models,
including gauge and fermionic fields.
\section{Gaussian Entropy from the Wigner Function}
\label{Gaussian entropy from the Wigner function}
Entropy has of course been widely studied over the years
\cite{Wehrl:1978zz, Mackey:1989zz, Barnett:1991zz,
Serafini:2003ke, Cacciatori:2008qs, Adesso:2007jg}. In this
section we will derive an entropy formula based on a na\"ive
probabilistic interpretation of the Wigner function. This
characterises much of the 20th century efforts to connect the
Wigner function, Boltzmann's equations, Boltzmann's H-theorem and
macroscopic entropy \cite{Hillery:1983ms}. Let us begin by
considering the Hamiltonian of a time dependent harmonic
oscillator, which is the one-particle ``free'' Hamiltonian:
\begin{equation}\label{Hamiltonian:1particle}
H(p,x) = \frac{p^{2}}{2m} + \frac{1}{2} m \omega^{2} x^{2} +
H_{\mathrm{s}} \,, \qquad H_{\mathrm{s}}= x j \,.
\end{equation}
In general, the oscillator's mass and frequency and the source
current can all depend on time: $m=m(t)$, $\omega=\omega(t)$ and
$j=j(t)$, respectively. $H_{\mathrm{s}}$ is the source Hamiltonian
and $j=j(t)$ the corresponding current. This Hamiltonian is
relevant both for many condensed matter systems that can be
realised in laboratories and in an early Universe setting. For the
latter, the time dependence of the parameters is introduced for
example by the Universe's expansion or by phase transitions.
Alternatively, this Hamiltonian can represent for example the
simplest model of a laser, where $x=x(t)$ is the photon amplitude,
where, if $m \rightarrow 1$, $\omega(t) \rightarrow E(t)/\hbar$ is
the photon's frequency and where $j=j(t)$ is a charge current that
coherently pumps energy into the system, transforming the vacuum
state into a coherent state \cite{Glauber:1963tx,Cahill:1969iq}.
The time dependence of the parameters changes the system's energy,
but simultaneously does not change the Gaussian nature of a state
if such a state is imposed initially, i.e.: the evolution implied
by the Hamiltonian~(\ref{Hamiltonian:1particle}) transforms an
initial Gaussian density matrix into another Gaussian density
matrix. Moreover, as we will come to appreciate, one has to add a
coupling to the environment or add a self-interaction to generate
entropy. The Lagrangian $L = p \dot{x} - H $ follows as usual from
the Hamiltonian. Note that the source term can be removed from
this Lagrangian by a simple coordinate shift:
\begin{equation}\label{coordinateshift}
x(t) \rightarrow z(t) = x(t) - x_{0}(t) \,,
\end{equation}
upon which the Lagrangian reduces to a quadratic form:
\begin{equation}\label{Lagrangian:quadraticform}
L = \frac{1}{2} m\dot{z}^{2} - \frac{1}{2}m\omega^{2}
z^{2}+L_{0}(t) + \frac{\mathrm{d}}{\mathrm{d}t}\left[m z
\dot{x}_{0} \right]\,.
\end{equation}
Here, $L_{0}(t)$ is a $z$-independent function of time and the
last term is a boundary term which does not contribute to the
equation of motion if $x_{0}(t)$ is a function that obeys:
\begin{equation}\label{shift:x0}
\frac{\mathrm{d}}{\mathrm{d}t} \left[m\dot{x}_{0}
\right]+m\omega^{2} x_{0} + j = 0 \,.
\end{equation}
A general Gaussian density matrix centered at the new origin
remains a Gaussian density matrix centered at the origin under the
evolution of a quadratic Hamiltonian and moreover, which is what
the analysis above shows, linear source terms in the Hamiltonian
will not alter the Gaussian nature of the state. We can thus
consider Gaussian density matrices centered at the origin, whose
time evolution is governed by an Hamiltonian of the
form~(\ref{Hamiltonian:1particle}) with $j(t) \rightarrow 0$. When
written in position space representation, the single particle
density operator of a quantum mechanical Gaussian state centered
at the origin is of the form:
\begin{subequations}
\label{density operator: particle}
\begin{equation} \label{density operator: particle1}
\hat{\rho}_{\mathrm{g}}(t) = \int_{-\infty}^{\infty} \mathrm{d}x
\int_{-\infty}^{\infty} \mathrm{d}y |x\rangle
\rho_{\mathrm{g}}(x,y;t)\langle y |\,,
\end{equation}
where:
\begin{equation} \label{density operator: particle2}
\rho_{\mathrm{g}}(x,y;t) = {\cal N}(t) \exp\left[
-a(t)x^2-b(t)y^2+2c(t)xy \right] \,,
\end{equation}
\end{subequations}
and where $a=a(t)$, $b=b(t)$ and $c=c(t)$ are determined from the
von Neumann equation~(\ref{density op:eom}). The subscript g
denotes ``Gaussian''. Moreover, from $\hat
\rho_{\mathrm{g}}^\dagger = \hat\rho_{\mathrm{g}}$ it follows that
$b^* = a$ and $c^*=c$. When $c=0$ one recovers a pure state with
vanishing entropy. However when $c\neq 0$ the density matrix is
mixed and entangled and it cannot be written in the simple
diagonal form, $\rho_{\mathrm{g}}(x,y;t) = \Psi^*(y,t) \Psi(x,t)$,
where $\Psi(x,t) = \sqrt{{\cal N}}\exp(-ax^2)$. When $c>0$ the
density matrix tends to get diagonal in the $x-y$ direction,
however when $c<0$ the diagonalisation occurs in the $x+y$
direction in which case the entropy is not defined. The Heisenberg
uncertainty relation is in trouble in this case as we will see in
equation (\ref{uncertainly relation}). Hence it is natural to
assume that $c>0$. We keep the discussion phenomenological and do
not discuss the physical origin of $c\neq 0$. The normalisation
${\cal N}$ is obtained from requiring
$\mathrm{Tr}[\hat{\rho}_{\mathrm{g}}]=1$:
\begin{equation}\label{Norm of rho}
\mathrm{Tr}[\hat{\rho}_{\mathrm{g}}] = \int_{-\infty}^{\infty}
\mathrm{d} \tilde{x} \langle \tilde x| \hat \rho_{\mathrm{g}} |
\tilde x \rangle = \int_{-\infty}^{\infty} \mathrm{d}x
\rho_{\mathrm{g}}(x,x;t) = {\cal N}
\sqrt{\frac{\pi}{2(a_{\mathrm{R}}-c)}}= 1 \,,
\end{equation}
from which we conclude:
\begin{equation}\label{Norm of rho:2}
{\cal N} = \sqrt{ \frac{2(a_{\mathrm{R}}-c)}{\pi}} \,,
\end{equation}
provided that $c < a_{\mathrm{R}}$ where $a_{\mathrm{R}}=\Re[a]$.
The equations that the functions $a(t)$, $b(t)$ and $c(t)$ of the
density matrix~(\ref{density operator: particle2}) obey can easily
be obtained from the von Neumann equation~(\ref{density op:eom}),
which in this Gaussian case in the amplitude basis reads:
\begin{equation} \label{density matrix:eom}
\imath\hbar\partial_t\rho_{\mathrm{g}}(x,y;t) = -
\frac{\hbar^2}{2m}\left(\partial^2_x-\partial^2_y\right)\rho_{\mathrm{g}}(x,y;t)
+ \frac12m\omega^2(x^2-y^2)\rho_{\mathrm{g}}(x,y;t) \,.
\end{equation}
If we insert equation (\ref{density operator: particle2}) in the
equation above we find:
\begin{subequations}
\label{density matrix:eom:2}
\begin{eqnarray}
\frac{\mathrm{d}a_{\mathrm{R}}}{\mathrm{d}t} &=&
\frac{4\hbar}{m}a_{\mathrm{I}} a_{\mathrm{R}} \label{density matrix:eom:2a}\\
\frac{\mathrm{d}c}{\mathrm{d}t} &=& \frac{4\hbar}{m}a_{\mathrm{I}}
c \label{density matrix:eom:2b}\\
\frac{\mathrm{d}}{\mathrm{d}t} \ln({\cal N}) &=&
\frac{2\hbar}{m}a_{\mathrm{I}} \label{density matrix:eom:2c}\\
\frac{\mathrm{d}a_{\mathrm{I}}}{\mathrm{d}t} &=&
\frac{2\hbar}{m}\left( a_{\mathrm{I}}^2 - a_{\mathrm{R}}^2 + c^2
\right) + \frac{m\omega^2}{2\hbar} \label{density matrix:eom:2d}
\,,
\end{eqnarray}
\end{subequations}
where we defined $a_{\mathrm{I}}=\Im[a]$. Note that ${\cal
N}\propto\sqrt{a_{\mathrm{R}}-c}$ is consistent with
equation~(\ref{density matrix:eom:2c}) as it
should\footnote{Indeed, by making use of
equations~(\ref{correlators:all}) one can show that
equations~(\ref{density matrix:eom:2}) are consistent with (and
equivalent to) the Hamilton equations for the
correlators~(\ref{Hamilton equations:correlators}).}.
The Wigner function is defined as a Wigner transform of the
density matrix $\rho_{\mathrm{g}}(x,y;t)$:
\begin{equation}\label{Wigner function:def}
{\cal W}(q,p;t) = \int_{-\infty}^{\infty} \mathrm{d} r {\rm
e}^{-\imath p r/\hbar}\rho_{\mathrm{g}}(q+r/2,q-r/2;t) \,,
\end{equation}
where we defined the average and relative coordinates, $q=(x+y)/2$
and $r = x-y$, respectively. Hence, a Wigner transform can be
thought of as an ordinary Fourier transform with respect to the
relative coordinate $r$ of the density matrix. The inverse Wigner
transform determines $\rho_{\mathrm{g}}(x,y;t)$ in terms of ${\cal
W}(q,p;t)$:
\begin{equation}\label{Wigner transform:inverse}
\rho_{\mathrm{g}}(x,y;t) = \int_{-\infty}^{\infty}
\frac{\mathrm{d} p}{2\pi\hbar}{\rm e}^{\imath p(x-y)/\hbar}{\cal
W}(q,p;t) \,,
\end{equation}
where $2\pi \hbar$ is the standard phase space measure. We obtain
the Gaussian Wigner function by performing the
integral~(\ref{Wigner function:def}):
\begin{equation}\label{Wigner function:Gauss}
{\cal W}(q,p;t) = {\cal M}(t) \exp\left[-\alpha_{\mathrm{w}}(t)
q^2 - \beta_{\mathrm{w}}(t) (p+q p_{\mathrm{c}}(t))^2 \right] \,,
\end{equation}
where:
\begin{subequations}
\label{Wigner function:Gauss2}
\begin{eqnarray}
\alpha_{\mathrm{w}} &=& 2(a_{\mathrm{R}}-c) \label{Wigner function:Gauss2a}\\
\beta_{\mathrm{w}} &=& \frac{1}{2\hbar^2 (a_{\mathrm{R}}+c)} \label{Wigner function:Gauss2b}\\
p_{\mathrm{c}}&=& 2\hbar a_{\mathrm{I}} \label{Wigner function:Gauss2c}\\
{\cal M} &=& \sqrt{\frac{4(a_{\mathrm{R}}-c)}{a_{\mathrm{R}}+c}}
\label{Wigner function:Gauss2d} \,,
\end{eqnarray}
\end{subequations}
and $p_{\mathrm{c}}$ denotes a ``classical'' momentum, which makes
the Wigner function non-diagonal. Note that the Wigner
function~(\ref{Wigner function:Gauss}) is normalised to unity, in
the sense that:
\begin{equation}\label{TrW=1}
\int_{-\infty}^{\infty} \frac{\mathrm{d}q\mathrm{d}p}{2\pi
\hbar}{\cal W}(q,p;t) = 1 \,.
\end{equation}
Recalling definition~(\ref{Wigner function:def}), equation
(\ref{density matrix:eom}) transforms to:
\begin{equation}
\Big(\partial_t + \frac{p}{m}\partial_q
- m\omega^2q\partial_p\Big) {\cal W}(q,p;t) = 0
\,, \label{Wigner function:eom}
\end{equation}
where we made use of the following relations:
\begin{subequations}
\label{intermediatestep1}
\begin{eqnarray}
\left(\partial_{x}^{2}-\partial_{y}^{2}\right)\rho_{\mathrm{g}}(x,y;t)
&=& 2
\partial_{q}\partial_{r} \rho_{\mathrm{g}}(q+r/2,q-r/2;t)
\label{intermediatestep1a}\\
r \mathrm{e}^{\imath p r/\hbar} &=& -\imath \hbar \partial_{p}
\mathrm{e}^{\imath p r /\hbar} \label{intermediatestep1b}\,.
\end{eqnarray}
\end{subequations}
Upon noting that $p/m = \dot x = \partial_p H$ and
$-m\omega^2x=\dot p =-\partial_x H$, we see that
equation~(\ref{Wigner function:eom}) is nothing but the Liouville
equation for the Boltzmann's distribution function $f(x,p;t)$,
\begin{equation}
\partial_t f(x,p;t) + \dot x \partial_x f(x,p;t) + \dot p \partial_p f(x,p;t) = 0
\label{Liouville}\,.
\end{equation}
A probabilistic interpretation of the Wigner function thus seems
plausible by the following identification:
\begin{equation}
{\cal W}(q,p;t) \Leftrightarrow f(x,p;t)
\,. \label{W=f}
\end{equation}
Of course, the equations for ${\cal W}(q,p;t)$ and $f(x,p;t)$ are
identical only in a free harmonic oscillator theory and, when
interactions are included, differences arise between the von
Neumann equation for the density operator $\hat
\rho_{\mathrm{g}}(t)$ (or the corresponding equation for ${\cal
W}(q,p;t)$) and the Boltzmann equation for $f(x,p;t)$. This indeed
makes the identification~(\ref{W=f}) less rigorous but it remains
a useful picture. The correspondence (\ref{W=f}) has been widely
used in the literature to develop a formalism relevant for example
for the physics of heavy ion collisions \cite{Mrowczynski:1992hq,
Aarts:2002dj} or for baryogenesis \cite{Prokopec:2003pj,
Prokopec:2004ic}.
In the distant 1969 Cahill and Glauber~\cite{Cahill:1969iq} have
proposed to approximate the density operator in the coherent state
basis $|\alpha\rangle$ by its diagonal form:
$\rho_{\mathrm{g}}\approx \rho_{\rm cs}(\alpha) = \langle
\alpha|\hat \rho_{\mathrm{g}}|\alpha\rangle$, which they interpret
as a quasi-probability distribution. Since coherent states harbour
many properties of classical particles, one can argue that this is
reasonable. Recall that coherent states are defined by $\hat
a|\alpha\rangle = \alpha|\alpha\rangle$, where $\hat a =
\sqrt{m\omega/2}[\hat x + \imath \hat p/(m\omega)]$ denotes the
annihilation operator for the
oscillator~(\ref{Hamiltonian:1particle}). The density matrix
$\rho_{\rm cs}(\alpha)$ represents an early example of a
quasi-probability distribution for a quantum system. Needless to
say, by this line of reasoning one can get only approximate
answers for the expectation value $\langle \hat A\rangle ={\rm
Tr}[\hat \rho_{\mathrm{g}}\hat A]$ of an operator $\hat A$ and
likewise for the entropy (which is just the expectation value of
$-\ln(\hat \rho_{\mathrm{g}})$). Another common way of
approximating expectation values of operators is to give a
probabilistic interpretation to the Wigner function, along the
same line of reasoning as argued above in equation~(\ref{W=f}) for
a free theory. In this approach~\cite{Cahill:1969iq,
Brandenberger:1992jh} the entropy is approximated by:
\begin{equation}
S \approx S_{\cal W} \equiv - {\rm Tr}[{\cal W}(q,p;t) \ln({\cal W}(q,p;t))]
\,, \label{entropy:Wigner}
\end{equation}
where the trace ${\rm Tr}\rightarrow \int
\mathrm{d}p\mathrm{d}q/[2\pi \hbar]$ should be interpreted as an
integration over the phase space volume as in
equation~(\ref{TrW=1}). We now insert~(\ref{Wigner
function:Gauss}) into~(\ref{entropy:Wigner}) to find:
\begin{eqnarray}
S_{\cal W} &=& \frac12 \ln\left(\frac{a_{\mathrm{R}}
+c}{a_{\mathrm{R}} -c}\right) + 1 -\ln(2)
\nonumber\\
&=& \ln\left(\frac{\Delta}{2}\right) + 1
\,, \label{Wigner function:entropy:final}
\end{eqnarray}
where we defined:
\begin{equation}
\Delta^2 = \frac{a_{\mathrm{R}} +c}{a_{\mathrm{R}} -c} \,.
\label{Delta function:2}
\end{equation}
We evaluated the phase space Gaussian integrals by shifting the
momentum integration to $p'=p-p_c$ (the Jacobian of this shift
equals unity). At this point, it is useful to evaluate a few
quantum mechanical expectation values. We can extract the
following correlators from our Gaussian state:
\begin{subequations}
\label{correlators:all}
\begin{eqnarray}
\langle \hat x^2\rangle &=& {\rm Tr}[\hat\rho_{\mathrm{g}}\hat
x^2] = \int_{-\infty}^{\infty} \mathrm{d}\tilde{x} \langle
\tilde{x}| \hat\rho_{\mathrm{g}}\hat x^2 | \tilde{x}\rangle =
\frac{1}{4(a_{\mathrm{R}} -c)}
\label{correlators:alla}\\
\Big\langle \frac12\{\hat x,\hat p\}\Big\rangle &=&
-\hbar\frac{a_{\mathrm{I}} }{2(a_{\mathrm{R}} -c)} \label{correlators:allb}\\
\langle \hat p^2\rangle &=& \hbar^2\frac{|a|^2-c^2}{a_{\mathrm{R}}
-c} \,. \label{correlators:allc}
\end{eqnarray}
\end{subequations}
Here, we made use of $\langle x| \hat p |\psi\rangle = -
\imath\hbar \partial_{x} \langle x|\psi\rangle$. Moreover, one can
easily verify that $ \langle [\hat{x},\hat{p}]\rangle = \langle
\hat x\hat p-\hat p\hat x\rangle = \imath\hbar $ as it should.
This enables us to find the equivalent inverse relations:
\begin{subequations}
\label{aI:aR:c}
\begin{eqnarray}
\label{aI:aR:ca} a_{\mathrm{I}} &=& -\frac{\left\langle
\frac12\{\hat x,\hat p\}\right\rangle}
{2\hbar\langle \hat x^2\rangle}
\label{aI:aR:cb}\\
a_{\mathrm{R}} &=& \frac{\Delta^2+1}{8\langle \hat x^2\rangle}
\label{aI:aR:cc}\\
c &=& \frac{\Delta^2-1}{8\langle \hat x^2\rangle} \,,
\label{aI:aR:cd}
\end{eqnarray}
\end{subequations}
and to express $\Delta(t)$ in equation (\ref{Delta function:2}) in
terms of the correlators~(\ref{correlators:all}):
\begin{equation}
\Delta^2(t) = \frac{4}{\hbar^2}
\left[
\langle \hat x^2\rangle \langle \hat p^2\rangle
- \Big\langle \frac12\{\hat x,\hat p\}\Big\rangle^2
\right]
\,. \label{Delta function:3}
\end{equation}
For the moment this is just a definition, but as we will come to
discuss, the physical meaning of $\hbar\Delta(t)$ is the phase
space area in units of $\hbar$ occupied by a (Gaussian) state
centered at the origin. For a pure state $\Delta(t) = 1$, while
for a mixed state $\Delta(t)>1$. Hence, we can define the
uncertainty relation for a general Gaussian state:
\begin{equation}
\frac{(\hbar\Delta(t))^2}{4} = \langle \hat x^2\rangle \langle \hat p^2\rangle
- \Big\langle \frac12\{\hat x,\hat p\}\Big\rangle^2
\geq \frac{\hbar^2}{4}
\,. \label{uncertainly relation}
\end{equation}
If $c$ is negative and $|c|< a_{\mathrm{R}}$, we can in principle
have $\Delta <1$, which can be seen from equation (\ref{Delta
function:2}). It would be interesting to investigate whether any
physical experiment could be proposed where a violation of
Heisenberg's uncertainty relation occurs, even if it were for only
a very brief moment in time (such that it would still hold on
average).
Furthermore, it is natural to define the statistical particle
number density on our phase space, in the quantum field
theoretical sense, as:
\begin{equation}
n(t) \equiv \frac{\Delta(t)-1}{2}
\,, \label{particle number}
\end{equation}
in terms of which the Wigner entropy~(\ref{Wigner
function:entropy:final}) can be rewritten as:
\begin{equation}
S_{\cal W}(t) = \ln\left(n(t)+\frac 12\right) + 1 \,.
\label{Wigner function:entropy:2}
\end{equation}
These results in essence agree with the entropy per mode of scalar
cosmological perturbations obtained in
\cite{Brandenberger:1992jh}. Shortly we shall see that the Wigner
entropy~(\ref{Wigner function:entropy:2}) is a good approximation
for the von Neumann entropy if $n\gg 1$, or equivalently when
$\hbar \Delta \gg 1$.
Let us try to put these results briefly in historical context. The
fundamental question that is left unanswered
in~\cite{Brandenberger:1992jh} is what the dynamical mechanism for
the growth of $\Delta$ is. Instead, the authors argue that an
observer that measures cosmological perturbations can be
approximated by a coherent state (which is a good classical
basis), thus explaining the growth in $\Delta$. Two mechanisms
were considered in \cite{Prokopec:1992ia}: a projection on the
particle number basis and on the coherent state basis. In both
cases, information of the phase of the squeezed Gaussian state is
lost, which thus generates entropy. In the latter approach
formula~(\ref{Delta function:3}) is reproduced, while in the
former, the entropy acquires the form of the statistical entropy
of Bose particles $S=(n+1)\ln(n+1)-n\ln(n)$ with $n\rightarrow
\sinh^2(r)$ the particle number associated with the squeezed
state, and $r$ the squeezing parameter of the state. While these
early articles, and many subsequent ones~\cite{Polarski:1995jg,
Kiefer:1999sj}, attempted to explain the origin of entropy of
cosmological perturbations by a (non-unitary) process of coarse
graining (associated with particular properties of the observer's
measuring apparatus), a genuine dynamical mechanism of entropy
generation was lacking.
This has changed recently \cite{Campo:2004sz, Campo:2005sy,
Prokopec:2006fc, Kiefer:2006je, Kiefer:2007zza, Campo:2008ju,
Campo:2008ij, Koksma:2009wa}, now dynamical mechanisms for entropy
generation have been proposed. Rather than discussing various
mechanisms in detail, let us just point out that the notion of how
entropy should be defined and generated has evolved: recently, it
has been advocated that the relevant entropy to consider is the
Gaussian entropy, although the approaches differ in the actual
origin of the entropy increase. For example, Prokopec and
Rigopoulos suggested in~\cite{Prokopec:2006fc} that tracing over
the unobserved isocurvature perturbations leads to entropy
generation in the adiabatic mode. Similarly, tracing over the
unobserved gravitational waves should lead to entropy generation
in the adiabatic mode~\cite{Koksma:2007zz}. Kiefer, Lohmar,
Polarski and Starobinsky~\cite{Kiefer:2006je,Kiefer:2007zza} as
well as Campo and Parentani~\cite{Campo:2008ju,Campo:2008ij},
among others, suggested that, if super-Hubble modes take the role
of the system, the origin of the entropy of cosmological
perturbations should be associated with the stochastic noise
generated by the sub-Hubble ultraviolet modes representing the
environment. Both Giraud and Serrau \cite{Giraud:2009tn} and we
\cite{Koksma:Talk,Koksma:Talk2,Koksma:2009wa} advocate that
neglecting observationally inaccessible non-Gaussianities is a
prominent mechanism to generate entropy.
\begin{figure}[t!]
\begin{minipage}[t]{.43\textwidth}
\begin{center}
\includegraphics[width=\textwidth]{GaussianPhaseSpace3alt.eps}
{\em \caption{General shape of the Wigner function,
representing the 2-dimensional Gaussian phase space distribution of a
certain state. The $1\sigma$ or $2\sigma$ cross-section is an
ellipse. For the current quantum mechanical discussion, we of course let
$\pi\rightarrow p$ and $\phi \rightarrow q$.
\label{fig:GaussianPhaseSpace3}}}
\end{center}
\end{minipage}
\vskip .1 cm
\begin{minipage}[t]{.43\textwidth}
\begin{center}
\includegraphics[width=\textwidth]{GaussianPhaseSpace1.eps}
{\em \caption{Squeezed state before decoherence. For example, the area of the ellipse of a vacuum state would be unity $\Delta=1$. For the current quantum mechanical discussion, we of course let
$\pi\rightarrow p$ and $\phi \rightarrow q$, also in figure \ref{fig:GaussianPhaseSpace2}. \label{fig:GaussianPhaseSpace1}}}
\end{center}
\end{minipage}
\hfill
\begin{minipage}[t]{.43\textwidth}
\begin{center}
\includegraphics[width=\textwidth]{GaussianPhaseSpace2.eps}
{\em \caption{Squeezed state after decoherence. Clearly, the
accessible phase space has increased and one can say that
knowledge about the system has been lost. The entropy has
increased compared to figure \ref{fig:GaussianPhaseSpace1}.
\label{fig:GaussianPhaseSpace2}}}
\end{center}
\end{minipage}
\end{figure}
So far, we have discussed how the Gaussian entropy in Wigner space
is defined and qualitatively discussed various mechanisms to
generate entropy dynamically. Let us now, to finish this section,
discuss how one can sensibly think of a decohered state in Wigner
space and discuss the physical meaning of $\Delta(t)$. Despite the
fact that the Wigner entropy (\ref{Wigner function:entropy:final})
only approximates the von Neumann entropy, it offers a useful,
intuitive way of depicting decohered states irrespective of the
precise underlying mechanism by which such a state has decohered
in the first place. Let us, for this reason, discuss this
phenomenologically. The Wigner function (\ref{Wigner
function:Gauss}) is a 2-dimensional Gaussian function whose width
in the $p$-direction need a priori not necessarily be the same as
in the $q$-direction. Hence, in general it can be squeezed in some
direction. This is illustrated in
figure~\ref{fig:GaussianPhaseSpace3}. The $1\sigma$ or $2\sigma$
cross-section of the phase space distribution has the shape of an
ellipse, which we also visualised in figure
\ref{fig:GaussianPhaseSpace3}. If the cross-section is a circle
centered at the origin, the state is either pure or mixed and if,
moreover, its area is unity $\Delta(t)=1$ it is a pure vacuum
state. If the cross-section is again a circle but displaced from
the origin, we are dealing with a generalised coherent-squeezed
state (whose shape is not described by equation (\ref{Wigner
function:entropy:final}) for obvious reasons). If the
cross-section is an ellipse, it is a squeezed state as mentioned
before. The ellipse is parametrised by:
\begin{equation}\label{Wignerslice}
\alpha_{\mathrm{w}}(t)q^{2}+\beta_{\mathrm{w}}(t)\left(p+q \,
p_{\mathrm{c}}(t)\right)^{2} = \vartheta\,,
\end{equation}
where we have made use of equation (\ref{Wigner function:Gauss})
and where $\vartheta$ is some constant that determines the height
at which we slice the Wigner function. The area of this ellipse is
now given by\footnote{Note that the area of ellipse defined by the
equation $Ax^{2}+Bxy+Cy^{2}=1$ is
$\mathcal{A}=2\pi/\sqrt{4AC-B^2}$.}:
\begin{equation}\label{Wignerarea}
\mathcal{A}(t) = \frac{\vartheta
\pi}{\sqrt{\alpha_{\mathrm{w}}(t)\beta_{\mathrm{w}}(t)}} =
\vartheta \pi \hbar \Delta(t) \,,
\end{equation}
where we have used equations (\ref{Wigner function:Gauss2}) and
(\ref{Delta function:2}). Clearly, the area of the ellipse in
Wigner space is determined straightforwardly from the phase space
area $\Delta(t)$. Note that if $\vartheta = 1/\pi$ we have
$\mathcal{A} = \Delta$.
Figures \ref{fig:GaussianPhaseSpace1} and
\ref{fig:GaussianPhaseSpace2} visualise the decoherence process of
a squeezed state. Initially, in figure
\ref{fig:GaussianPhaseSpace1}, the state is squeezed and has unity
phase space area $\Delta(t_{0})=1$, such that its entropy
vanishes. Now we switch on our favourite decoherence mechanism:
the state interacts with some environment and environmental
degrees of freedom are traced over or non-Gaussian correlators
generated by the interaction with the environment are neglected.
Consequently, the area in phase space increases $\Delta(t)>1$,
such that $S(t)>0$. Hence, the ellipse in Wigner space grows which
we depict in figure \ref{fig:GaussianPhaseSpace2}. An important
feature of any decoherence process reveals itself: for a highly
squeezed state, given the knowledge of $q$ after some measurement,
the value $p$ can take in a subsequent measurement is constraint
in a very narrow interval. Interestingly, after the state has
decohered, and given the same measurement of $q$, the range of
values $p$ can take has increased. Indeed one can say that
knowledge about $p$ has been lost compared to the squeezed state
before decoherence, hence entropy has been generated.
Alternatively, we can say that given some measurement of $q$, the
value $p$ can take is, after decoherence, drawn from a classical
stochastic distribution, which indeed coincides with the familiar
idea that decohered quantum systems should behave as uncorrelated
stochastic systems. Indeed, $n$ roughly counts the number of
patches in phase space that behave independently and are
uncorrelated.
With this intuitive notion of decohered state in mind, one can
understand that, for highly squeezed states, the position space
basis (which generalises to the field amplitude basis in the
quantum field theoretical case) is the pointer basis. Let us start
out with a highly squeezed state, i.e.: take for example a pure
state during inflation that rapidly squeezes during the Universe's
expansion. The Hamiltonian of the state is then dominated by the
potential term, and the kinetic term only contributes little. In
thermal equilibrium, the system will minimise its free energy
$F=H-TS$, where $T$ is the temperature of the heat bath. Now, if
we switch on some decoherence mechanism due to interaction with
the environment, the entropy will increase mainly due to momentum
increase, i.e.: $\langle \hat{p}^2 \rangle$ will increase. Indeed,
increasing $\langle \hat{p}^2 \rangle$ will hardly affect the
Hamiltonian, whereas it will significantly affect the entropy.
Hence, the variable $x$ is robust during the process of
decoherence, in the sense that $\langle \hat{x}^{2} \rangle$ will
hardly change such that it qualifies as a proper pointer basis.
Note however, that $x$ is only a pointer basis in the statistical
sense, such that there is a well defined probability distribution
function from which a measurement can be drawn.
\section{Gaussian entropy from the replica trick}
\label{Gaussian entropy from the replica trick}
The entropy of a general Gaussian state has been derived by Sohma,
Holevo and
Hirota~\cite{SohmaHolevoHirota:1999,SohmaHolevoHirota:2000} by
making use of Glauber's P representation for the density operator.
Here we present an alternative derivation for the von Neumann
entropy~(\ref{entropy:vN}) of a general Gaussian state by making
use of the replica trick (see e.g.:
\cite{Callan:1994py,Calabrese:2009qy,Calabrese:2004eu}). Of
course, there are other analogous methods one can use\footnote{We
thank Theo Ruijgrok for his useful comments.}. Notice first that:
\begin{equation}
\ln(\hat\rho_{\mathrm{g}}) = \lim_{\epsilon\rightarrow 0}
\frac{\hat\rho_{\mathrm{g}}^\epsilon-1}{\epsilon}
\,.
\label{replica trick}
\end{equation}
Here, $\epsilon$ is a positive integer which is analytically
extended to zero. Hence, in order to calculate the
entropy~(\ref{entropy:vN}) we need to evaluate ${\rm
Tr}\left[\hat\rho_{\mathrm{g}}^{1+\epsilon}\right]$. Using
equation (\ref{density operator: particle}), we thus have:
\begin{eqnarray}
{\rm Tr}\left[\hat\rho_{\mathrm{g}}^{1+\epsilon}\right]
&=& {\cal N}^{1+\epsilon}
\int_{-\infty}^{\infty} \mathrm{d}x_0 \cdots \int_{-\infty}^{\infty} \mathrm{d}x_{\epsilon}\exp \left[-2a_{\mathrm{R}} \left(x_0^2+\cdots +x_\epsilon^2 \right)
+2c \left(x_0x_1 + x_1x_2 + \cdots + x_\epsilon x_0 \right) \right]
\nonumber\\
&=& {\cal N}^{1+\epsilon}\int_{-\infty}^{\infty} \mathrm{d}x_0 \cdots \int_{-\infty}^{\infty} \mathrm{d}x_{\epsilon}\exp
\left[-2\beta \left\{ (x_0 - \alpha x_1)^2+(x_1 - \alpha x_2)^2+ \cdots +
(x_\epsilon - \alpha x_0)^2 \right\} \right]
\nonumber\\
&=& |1-\alpha^{1+\epsilon}|^{-1}
\left({\cal N}\int_{-\infty}^{\infty} \mathrm{d}y \exp\left[ -2\beta y^2 \right] \right)^{1+\epsilon}
\nonumber\\
&=& |1-\alpha^{1+\epsilon}|^{-1}
\left(\frac{a_{\mathrm{R}}-c}{\beta}\right)^{\frac{1+\epsilon}{2}}\,.
\label{replica trick:2}
\end{eqnarray}
Here, $J=|1-\alpha^{1+\epsilon}|^{-1}$ is the Jacobian of the
transformation from $x_i$ to $y_i=x_i-\alpha x_{i+1}$, where
$i=\{0,1,\cdots ,\epsilon-1\}$, and $y_\epsilon=x_\epsilon-\alpha
x_{0}$. A comparison of the first and the second line
in~(\ref{replica trick:2}) tells us that:
\begin{subequations}
\label{replica trick:3}
\begin{eqnarray}
\alpha_\pm &=&
\frac{a_{\mathrm{R}}}{c}\pm\sqrt{\left(\frac{a_{\mathrm{R}}}{c}\right)^2-1}
\label{replica trick:3a}\\
\beta_\pm &=& \frac{c}{2\alpha_\pm} = \frac{c\alpha_\mp}{2} \,.
\label{replica trick:3b}
\end{eqnarray}
\end{subequations}
In the limit $c\rightarrow 0$ one must recover a pure diagonal
density operator and, consequently, the shift $\alpha$ should
vanish, singling out:
\begin{subequations}
\label{replica trick:4}
\begin{eqnarray}
\alpha &=& \alpha_- = \frac{a_{\mathrm{R}}}{c} -
\sqrt{\left(\frac{a_{\mathrm{R}}}{c}\right)^2-1}
\label{replica trick:4a}\\
\beta &=& \beta_- \,. \label{replica trick:4b}
\end{eqnarray}
\end{subequations}
in equation~(\ref{replica trick:3}) as the physical choice. Based
on $(a_{\mathrm{R}}-c)/\beta=(1-\alpha)^2$, we can rewrite
equation~(\ref{replica trick:2}) in a simpler form:
\begin{equation}
{\rm Tr}\left[\hat\rho_{\mathrm{g}}^{1+\epsilon}\right]
= \frac{(1-\alpha)^{1+\epsilon}}{|1-\alpha^{1+\epsilon}|}
\,. \label{replica trick:5}
\end{equation}
Recall that normalisability of the density operator requires
$a_{\mathrm{R}}>c$, implying that $0\leq\alpha<1$, such that the
absolute value in (\ref{replica trick:5}) can be dropped. This
means that ${\rm
Tr}\left[\hat\rho_{\mathrm{g}}^{1+\epsilon}\right]$ is analytic in
$\alpha$ (for positive integers $\epsilon$):
\begin{equation}
{\rm Tr}\left[\hat\rho_{\mathrm{g}}^{1+\epsilon}\right]
= \frac{(1-\alpha)^{1+\epsilon}}{1-\alpha^{1+\epsilon}}
\,, \label{replica trick:6}
\end{equation}
such that one can analytically extend to complex $\epsilon$.
According to the replica method, the entropy is then obtained by
taking the limit $\epsilon\rightarrow 0$. To perform this
procedure we expand~(\ref{replica trick:6}) to linear order in
$\epsilon$ and we find the following expression for the von
Neumann entropy~(\ref{entropy:vN}):
\begin{equation}
S_{\rm vN}(t) = -\ln(1-\alpha)-\frac{\alpha}{1-\alpha}\ln(\alpha)
\,. \label{entropy:rho log rho:final}
\end{equation}
Using the expression for $\alpha$ in equation~(\ref{replica
trick:4a}) and the definition of the phase space area in equations
(\ref{Delta function:2}) and~(\ref{Delta function:3}), we find:
\begin{equation}
\alpha(t) = \frac{\Delta(t)-1}{\Delta(t)+1} \,,
\label{alphatoDelta}
\end{equation}
such that we can express the entropy~(\ref{entropy:rho log
rho:final}) solely in terms of $\Delta$:
\begin{equation}
S_{\rm vN}(t) =
\frac{\Delta+1}{2}\ln\left(\frac{\Delta+1}{2}\right) -
\frac{\Delta-1}{2}\ln\left(\frac{\Delta-1}{2}\right)\,.
\label{entropy:final}
\end{equation}
Since $0\leq\alpha(t)<1$, we have $1 \leq \Delta(t) < \infty$.
The von Neumann entropy~(\ref{entropy:final}) vanishes when
$\Delta=1$ (or $c=0$), defining a pure state, while a strictly
positive entropy $S_{\rm vN}>0$ implies a mixed state $\Delta >
1$. Clearly, the Gaussian von Neumann entropy is completely
determined by the three Gaussian correlators characterising the
state (\ref{Delta function:3}). Relations~(\ref{entropy:final})
and~(\ref{entropy:rho log rho:final}) suggest the following
definition of the particle number $n(t)$:
\begin{equation}
n(t) \equiv \frac{\alpha(t)}{1-\alpha(t)} = \frac{\Delta(t) -1}{2}
\,, \label{n}
\end{equation}
where $0\leq n<\infty$, in terms of which the
entropy~(\ref{entropy:rho log rho:final}), (\ref{entropy:final})
becomes:
\begin{equation}
S_{\rm vN}(t)=(1+n)\ln(1+n)-n\ln(n) \,, \label{entropy:n}
\end{equation}
which is the well known result from statistical physics for the
entropy of $n$ Bose particles per (quantum)
state~\cite{Kubo:1965}. Note that (\ref{entropy:n}) is a convex
function of $n$, such that $S(n_1)+S(n_2)
> S(n_1+n_2)$, which is another desirable property for the entropy. The particle number $n$ defined
in~(\ref{n}) should be interpreted as the number of independent
(uncorrelated) regions in the phase space of a single (one
particle) quantum state.
We will show next that $\Delta$ is a special function, as it is
conserved by the evolution equations resulting from the
Hamiltonian~(\ref{Hamiltonian:1particle}):
\begin{subequations}
\label{Hamilton equations}
\begin{eqnarray}
\dot{\hat x} &=& \partial_p \hat H = \frac{\hat p}{m}
\label{Hamilton equationsa} \\
\dot{\hat p} &=& -\partial_x \hat H = - m\omega^2\hat x \,,
\label{Hamilton equationsb}
\end{eqnarray}
\end{subequations}
where we set $j\rightarrow 0$. These equations imply for the
correlators:
\begin{subequations}
\label{Hamilton equations:correlators}
\begin{eqnarray}
\frac{\mathrm{d}}{\mathrm{d}t}\langle\hat x^2\rangle
&=& \frac{2}{m}\Big\langle\frac12\{\hat x,\hat p\}\Big\rangle
\label{Hamilton equations:correlatorsa}\\
\frac{\mathrm{d}}{\mathrm{d}t}\Big\langle\frac12\{\hat x,\hat
p\}\Big\rangle
&=& - m\omega^2\langle\hat x^2\rangle
+ \frac{1}{m}\langle\hat p^2\rangle
\label{Hamilton equations:correlatorsb}\\
\frac{\mathrm{d}}{\mathrm{d}t}\langle[\hat x,\hat p]\rangle &=& 0
\label{Hamilton equations:correlatorsc}\\
\frac{\mathrm{d}}{\mathrm{d}t}\langle\hat p^2\rangle
&=& - 2m\omega^2\Big\langle\frac12\{\hat x,\hat p\}\Big\rangle
\label{Hamilton equations:correlatorsd}\,.
\end{eqnarray}
\end{subequations}
Recall that these equations are, in the light of
equation~(\ref{aI:aR:c}), equivalent to~(\ref{density
matrix:eom:2}). The third equation is automatically satisfied by
the commutation relations, $[\hat x,\hat p] = \imath \hbar$. One
can combine the other three equations in~(\ref{Hamilton
equations:correlators}) to show that:
\begin{equation}
\frac{\mathrm{d}}{\mathrm{d}t} \frac{\hbar^2\Delta^2}{4}
= \frac{\mathrm{d}}{\mathrm{d}t}\left[\langle\hat x^2\rangle \langle\hat p^2\rangle
- \Big\langle\frac12\{\hat x,\hat p\}\Big\rangle^2\right]
= 0
\label{Delta:conservation}
\end{equation}
This implies that $\Delta$ (and in fact any function of $\Delta$)
is conserved by the Hamiltonian evolution~(\ref{Hamilton
equations}--\ref{Hamilton equations:correlators}).
Of course, when interactions are included, the conservation law
(\ref{Delta:conservation}) becomes approximate. In quantum field
theory interactions are typically cubic or quartic in the fields
and will thus induce non-Gaussianities in the density matrix. In
particular, for a quantum mechanical $\lambda\phi^4$ model, this
has been investigated by Calzetta and Hu \cite{Calzetta:2002ub,
Calzetta:2003dk}. They derive an H-theorem for $\Delta$ in the
case of a quantum mechanical $O(N)$ model.
\begin{figure}[t!]
\begin{minipage}[t]{.43\textwidth}
\begin{center}
\includegraphics[width=\textwidth]{Entropy1.eps}
\end{center}
\end{minipage}
\hfill
\begin{minipage}[t]{.43\textwidth}
\begin{center}
\includegraphics[width=\textwidth]{Entropy2.eps}
\end{center}
\end{minipage}
\begin{minipage}{.86\textwidth}
\begin{center}
{\em \caption{Left: The von Neumann entropy (solid line) and its
Wigner entropy approximation (dashed line) for a Gaussian state as
a function of the phase space area $\Delta$. Right: The difference
between the Wigner and the von Neumann entropy. The Wigner entropy
approaches the exact von Neumann entropy rapidly as $\Delta$
grows. \label{fig:entropycomp}}}
\end{center}
\end{minipage}
\end{figure}
Finally, it is interesting to compare the von Neumann
entropy~(\ref{entropy:final}) and the Wigner entropy~(\ref{Wigner
function:entropy:final}), as shown in
figure~\ref{fig:entropycomp}. This reveals that the Wigner entropy
represents a large phase space (semiclassical) limit of the von
Neumann entropy, which should not come as a surprise. Indeed, the
Wigner entropy can be used to accurately represent the entropy in
systems which develop large correlators (occupation numbers), such
as cosmological perturbations~\cite{Brandenberger:1992jh}. The
Wigner entropy fails however to take account of quantum
correlations present in the density matrix and Wigner function,
which are properly taken into account in the exact expression for
the entropy of a general Gaussian state~(\ref{entropy:final}).
\section{Non-Gaussian entropy: Two examples}
\label{Non-Gaussian entropy: Two examples}
An important question is how to generalise the von Neumann entropy
of a Gaussian state to include non-Gaussian corrections.
Non-Gaussianities can be important either because they are made
large on purpose by specially preparing the state, or simply
because they are accessible in measurements due to the fact that
the observer's measurement device is very accurate.
While we are far from building a general theory of how
non-Gaussian correlations affect the notion of phase space volume,
statistical particle number and thus entropy, we shall present two
examples in this section: we consider a system in a state which
can be represented by an admixture of a Gaussian density matrix
and some small non-Gaussian contributions. Secondly, we consider
small quartic corrections to the density matrix, contributing to
the kurtosis of the ground state.
Qualitatively, we expect that the correction to the Gaussian
entropy due to non-Gaussianities present in a theory is small,
except for systems whose entropy is (almost) zero. This can most
easily be appreciated in the Wigner approach to entropy. Consider
for example a system with a large entropy, for example a squeezed
state whose phase space area in Wigner space has increased
significantly due to developing a $\Delta \gg 1$.
Non-Gaussianities can deform the area in phase space of the state
(by squeezing, pushing or stretching the Wigner function) roughly
by a factor of order unity. Consequently, the effect on the
entropy will be relatively small. This can be understood by
realising that a lot of information is contained in unequal time
correlators. Of course, such an argument ceases to be true for a
pure state, whose quantum properties are very pronounced.
\subsection{Admixture of the Ground State and First Excited State} \label{Admixture of the Ground State and First Excited State}
Let us consider a density matrix of the form:
\begin{equation}
\rho_2(x,y;t) = {\cal N}_2(1+ \zeta_0 x + \zeta_0^*y+\zeta_1 x^2 + \zeta_1^*y^2+2\zeta_2 xy)
\exp\left(-ax^2-a^*y^2+2cxy\right)
\,. \label{rho2}
\end{equation}
Such a state typically appears when one uses a laser to excite an
harmonic oscillator in its ground state. The system can then be
described by an admixture of the ground state and a small
contribution to the first exited state\footnote{Recall that the
wave function of the pure one particle state of the simple
harmonic oscillator~(\ref{Hamiltonian:1particle}) (with $j=0$) is:
\begin{equation}
\psi_1(x) = \langle x|\psi_1 \rangle =
\bigg[\frac{4}{\pi}\bigg(\frac{m\omega}{\hbar}\bigg)^3\bigg]^{1/4}
\, x\, \exp\Big(-\frac{m\omega x^2}{2\hbar}\Big)
\,, \nonumber
\end{equation}
with energy $E_1 = \hbar \omega\left(1+\frac12\right)$. The
problem at hand reduces to this state upon identifying
$a_{\mathrm{R}} \leftrightarrow m\omega/(2\hbar)$,
$a_{\mathrm{I}}\rightarrow 0$, $c\rightarrow 0$,
$\zeta_0\rightarrow 0$ and $\zeta_1\rightarrow 0$.}. Note that the
process of pumping energy in a simple harmonic oscillator in its
ground state could lead to creating a state that is not pure ($c
\neq 0$). The measuring apparatus is assumed to be sensitive to
the admixture of the two states. We shall assume that
non-Gaussianity is small, in the sense that all $\zeta_i$
parameters ($i=0,1,2$) are small, such that we content ourselves
with performing the analysis up to linear order in $\zeta_i$.
Notice first that to this order the terms in parentheses in
equation~(\ref{rho2}) can be approximated by an exponential:
\begin{equation}
1+ \zeta_0 x + \zeta_0^*y+\zeta_1 x^2 + \zeta_1^*y^2+2\zeta_2 xy
\simeq \exp\left(\zeta_0 x + \zeta_0^*y+\zeta_1 x^2 +
\zeta_1^*y^2+2\zeta_2 xy\right) \,, \label{exponentialassumption}
\end{equation}
implying that the first two terms $\zeta_0 x+\zeta_0^*y$ induce an
entropy conserving shift in $x$ and $y$ (as in coherent states).
Of course, this ceases to be true at quadratic and higher orders
in $\zeta_0$. On the other hand, the $\zeta_1$ and $\zeta_2$ terms
do change the entropy even at linear order. At higher order,
$\zeta_1$ and $\zeta_2$ induce kurtosis whose effect on the
entropy we will discuss shortly.
Before we turn our attention to calculating the entropy, let us
first show that a density matrix (\ref{rho2}) with $\zeta_0
\rightarrow 0$ indeed generates non-trivial higher order
correlators. Normalising the trace to unity yields:
\begin{equation}
{\cal N}_2 = \sqrt{\frac{2}{\pi}}
\frac{(a_{R}-c)^{\frac{3}{2}}}{a_{\mathrm{R}}+c+\frac{1}{2}(\zeta_{1R}+\zeta_2)}
\,. \label{normdensity2}
\end{equation}
An interesting higher order correlator to consider is for example
the connected part of the four-point correlator:
\begin{equation}
\langle \hat x^{4} \rangle_{\mathrm{con}} = \langle \hat x^{4}
\rangle - 3(\langle \hat x^{2} \rangle)^{2} \,.
\label{4pntcorrelator}
\end{equation}
If the connected part of the four-point correlator is non-zero,
the state is said to have kurtosis. Using the Gaussian density
matrix in equation (\ref{density operator: particle}) one can
easily show that the connected part of the four-point correlator
vanishes as it should: for free theories higher order correlators
either vanish or can be expressed in terms of Gaussian
correlators. Using however the non-Gaussian density matrix in
equation (\ref{rho2}) with $\zeta_0 \rightarrow 0$ and the
normalisation constant (\ref{normdensity2}) we find:
\begin{equation}
\langle \hat x^{4} \rangle_{\mathrm{con}} = -
\frac{3(\zeta_{1R}+\zeta_2)^{2}}{16(a_{\mathrm{R}}-c)^{2}(a_{\mathrm{R}}-c+\frac{1}{2}(\zeta_{1R}+\zeta_2))^{2}}
\simeq -
\frac{3(\zeta_{1R}+\zeta_2)^{2}}{16(a_{\mathrm{R}}-c)^{4}}
+\mathcal{O}\left((\zeta_{1R})^{3},(\zeta_2)^{3}\right) \,.
\label{4pntcorrelator2}
\end{equation}
Clearly, genuine non-Gaussian correlators are generated and this
state has kurtosis at quadratic order.
Let us now calculate the entropy by making use of equation
(\ref{exponentialassumption}). As said, it is precisely
$\zeta_{1}$ and $\zeta_{2}$ that affect the entropy. To see that
notice further that their effect can be captured by a shift in $a$
and $c$ as follows:
\begin{subequations}
\label{a+c:rescaling}
\begin{eqnarray}
a &\rightarrow & \bar a = a - \zeta_1 \label{a+c:rescalinga}\\
c &\rightarrow & \bar c = c + \zeta_2 \label{a+c:rescalingb}\,,
\end{eqnarray}
\end{subequations}
such that the corresponding von Neumann entropy simply reads ({\it
cf.} equation (\ref{entropy:n})):
\begin{equation}
S_2 = (1+\bar n)\ln(1+\bar n)-\bar n\ln(\bar n) \label{s2}
\,,
\end{equation}
where $\bar n = n + \delta n$ and where as usual $n=(\Delta
-1)/2$. Now $\delta n\propto \zeta_{1R},\zeta_2$ is the correction
to the Gaussian entropy we are about to calculate. Comparing with
equations~(\ref{replica trick:3}) and~(\ref{n}) we have:
\begin{subequations}
\label{nandalphabar}
\begin{eqnarray}
\bar \alpha &=& \frac{\bar a_{\mathrm{R}}}{\bar c} -
\sqrt{\frac{\bar
a_{\mathrm{R}}^2}{\bar c^2}-1} \label{nandalphabara}\\
\bar n &=& \frac{\bar \alpha}{1-\bar\alpha}
\label{nandalphabarb}\,,
\end{eqnarray}
\end{subequations}
from which we immediately find, to linear order in $\zeta_{1R}$
and $\zeta_{2}$:
\begin{subequations}
\label{deltananddeltaalpha}
\begin{eqnarray}
\delta\alpha &=& \frac{\alpha(1-\alpha)}{1+\alpha}
\left(\frac{\zeta_{1R}}{a_{\mathrm{R}}-c}+
\frac{1+\alpha^2}{2\alpha}\frac{\zeta_2}{a_{\mathrm{R}}-c}\right)
\label{deltananddeltaalphaa} \\
\delta n &=& \frac{\delta\alpha}{(1-\alpha)^2}
= \frac{n(1+n)}{1+2n}
\left(\frac{\zeta_{1R}}{a_{\mathrm{R}}-c}+
\frac{1+2n+2n^2}{2n(1+n)}\frac{\zeta_2}{a_{\mathrm{R}}-c}\right)
\label{deltananddeltaalphab}\,,
\end{eqnarray}
\end{subequations}
where we used $a_{\mathrm{R}}/c = (1+\alpha^2)/(2\alpha)$.
Finally, $\delta S_{2} = \{\ln[(1+n)/n]\}\delta n$ implies that:
\begin{subequations}
\label{s2:final}
\begin{equation}
S_2(t)=S_{\mathrm{g}}(t) + \delta S_{2}(t) \label{s2:finala} \,,
\end{equation}
where $\delta S_{2}$ is an entropy shift given by:
\begin{equation}
\delta S_2 (t)= \frac{n(1+n)}{1+2n}
\left(\frac{\zeta_{1R}}{a_{\mathrm{R}}-c}+
\frac{1+2n+2n^2}{2n(1+n)}\frac{\zeta_2 }{a_{\mathrm{R}}-c}\right)
\ln\left(\frac{1+n}{n}\right) \label{s2:finalb}\,,
\end{equation}
\end{subequations}
where $S_{\rm g}=(1+n)\ln(1+n)-n\ln(n)$ is the von Neumann entropy
of a Gaussian state and $n$ is the statistical particle number
associated with the Gaussian part of the state as before.
Several comments are in order. Firstly, the
result~(\ref{s2:final}) implies that to linear order
$\Im[\zeta_1]$ does not change the entropy. In fact,
$\Im[\zeta_1]$ induces (de)squeezing of the state, which can be
appreciated from equation (\ref{Wigner function:Gauss2c}).
Secondly, $\delta S_2$ is positive (negative) whenever
$\zeta_{1R}$ and $\zeta_2$ are positive (negative), irrespective
of $n>0$. Thirdly, we have rescaled $\zeta_{1R}$ and $\zeta_2$ by
$(a_{\mathrm{R}}-c)$ to get a dimensionless quantity. This
rescaling is natural, since $a_{\mathrm{R}}-c$ measures the width
of the state. Finally, for large $n$ the formula~(\ref{s2:final})
gives a meaningful answer, since:
\begin{equation}
\lim_{n\rightarrow \infty} \delta S_2 =
\frac{1}{2}\frac{\zeta_{1R}+\zeta_2}{a_{\mathrm{R}}-c}\,,
\label{Entropylimit1}
\end{equation}
is finite. On the other hand, when $n\rightarrow 0$, we encounter
a logarithmic divergence:
\begin{equation}
\lim_{n\rightarrow 0} \delta S_2 =
\left(n\frac{\zeta_{1R}}{a_{\mathrm{R}}-c}+
\frac{1}{2}\frac{\zeta_2}{a_{\mathrm{R}}-c}\right)
\ln\Big(\frac{1}{n}\Big)\,, \label{Entropylimit2}
\end{equation}
indicating a mild (logarithmic) breakdown of the linear expansion.
Notice however that also in that limit the formula~(\ref{s2})
remains applicable.
\subsection{Kurtosis}
\label{Kurtosis}
In the former example non-Gaussianities are generated by adding a
quadratic polynomial to the prefactor. One could also introduce
non-Gaussianities by adding higher order, i.e.: cubic or quartic,
powers in the exponential. Cubic corrections generate skewness and
since they contribute to the entropy at quadratic and higher
orders in the skewness parameter, we shall now focus on the
quartic corrections to the density matrix. Quartic corrections to
a density matrix affect the kurtosis of a state and we parametrise
this as follows:
\begin{equation}
\rho_4(x,y;t) = {\cal N}_4\exp \left(-ax^2-a^*y^2+2cxy \right)\exp \left(\eta_0x^4+\eta_0^*y^4+\eta_1x^3y+\eta_1^*xy^3+\eta_2x^2y^2 \right)
\equiv \frac{{\cal N}_4}{{\cal N}_{\rm g}} \rho_{\rm g} (x,y;t) \rho_{\rm
ng}(x,y;t),
\label{rho4}
\end{equation}
where $\rho_{\rm g}$ is the Gaussian density matrix (\ref{density
operator: particle}) as before, with the normalisation ${\cal
N}_{\rm g}$ given in equation (\ref{Norm of rho:2}), and where
finally $\rho_{\rm
ng}=\exp(\eta_0x^4+\eta_0^*y^4+\eta_1x^3y+\eta_1^*xy^3+\eta_2x^2y^2)$.
As in the previous example, we shall consider only linear
corrections in $\eta_i$ ($i=0,1,2$) to the entropy. To this order
the non-Gaussian part of~(\ref{rho4}) can also be written as:
\begin{equation}
\delta \rho_{\rm ng} \equiv \rho_{\rm ng} - 1 =
\eta_0x^4+\eta_0^*y^4+\eta_1x^3y+\eta_1^*xy^3+\eta_2x^2y^2 +{\cal
O}(\eta_i^2) \,. \label{rhong:lin}
\end{equation}
We can find the normalisation constant after some simple algebra
by making use of~(\ref{rhong:lin}):
\begin{equation}
{\cal N}_{4} = {{\cal N}_{\rm g}}
\bigg[1+\frac34 \frac{\eta_0+\eta_0^*+\eta_1+\eta_1^*+\eta_2}{[2(a_{\mathrm{R}}-c)]^2}
\bigg]^{-1}
\,. \label{calN4}
\end{equation}
\begin{figure}[t!]
\begin{minipage}[t]{.43\textwidth}
\begin{center}
\includegraphics[width=\textwidth]{GaussianPhaseSpace4.eps}
\end{center}
\end{minipage}
\hfill
\begin{minipage}[t]{.43\textwidth}
\begin{center}
\includegraphics[width=\textwidth]{GaussianPhaseSpace5.eps}
\end{center}
\end{minipage}
\begin{minipage}{.86\textwidth}
\begin{center}
{\em \caption{Wigner transform of the density matrix described
by equations (\ref{rho4}) and (\ref{rhong:lin}), with $\eta_0
\neq 0$ only. We used $\eta_0>0$ (left) and $\eta_0<0$ (right). The peak structure of our state in phase space, that was so simple for Gaussian
density matrices, is much more complicated. Some regions in
phase space have a negative value of the
Wigner function, indicating a breakdown of the use of a Wigner
function as a convenient measure of probability.
\label{fig:NGWigner}}}
\end{center}
\end{minipage}
\end{figure}
To gain some intuitive understanding of what our density matrix
looks like, let us consider figure \ref{fig:NGWigner}. Here, we
show the Wigner transform of equation (\ref{rho4}) using
(\ref{rhong:lin}), with $\eta_1 = 0 = \eta_2$. Clearly, the
quartic corrections change the peak structure of our state in
phase space. Moreover, some regions in phase space now have a
negative Wigner function. This nicely illustrates the limitations
of using the Wigner function as a probability density on our phase
space as soon as non-Gaussianities (due to interactions) are
included, i.e.: equation (\ref{W=f}) holds only approximately in
this case.
In order to calculate the entropy, we need ${\rm Tr}[\hat
\rho_4^{1+\epsilon}]$ as before ({\it cf.} equation~(\ref{replica
trick:2})). To linear order in $\delta \rho_{\rm ng}$ we have:
\begin{equation}
{\rm Tr}[\hat \rho_4^{1+\epsilon}] = \left(\frac{{\cal N}_4}{{\cal
N}_{\rm g}}\right)^{1+\epsilon}
\Big( {\rm Tr}[\hat \rho_{\rm g}^{1+\epsilon}]
+ (1+\epsilon){\rm Tr}[\hat \rho_{\rm g}^{1+\epsilon}\delta\hat\rho_{\rm ng}]
\Big)
\,. \label{rho4:lin}
\end{equation}
The first term in the parentheses is just the Gaussian
result~(\ref{replica trick:2}) while the latter term can be
evaluated by making use of the formulae in Appendix \ref{Appendix
A Useful integrals}. Expanding to linear order in $\epsilon$, the
equation above yields:
\begin{eqnarray}
{\rm Tr}[\hat \rho_4^{1+\epsilon}]
&=& 1 + \epsilon\Big[\ln(1-\alpha)+\frac{\alpha}{1-\alpha}\ln(\alpha)\Big]
+\epsilon
\bigg[
\frac34\frac{\eta_0+\eta_0^*}{[2(a_{\mathrm{R}}-c)]^2}
\Big(\frac{4\alpha}{1-\alpha^2}\ln(\alpha)-1\Big)
\nonumber\\
&&\hskip 1cm
+\,\frac34\frac{\eta_1+\eta_1^*}{[2(a_{\mathrm{R}}-c)]^2}
\Big(\frac{1+\alpha}{1-\alpha}\ln(\alpha)-1\Big)
+\frac34\frac{\eta_2}{[2(a_{\mathrm{R}}-c)]^2}
\Big(\frac43\frac{1+\alpha+\alpha^2}{1-\alpha^2}\ln(\alpha)-1\Big)
\bigg]
\,.\qquad \label{rho4:lin:b}
\end{eqnarray}
As expected, we see that the entropy naturally splits into a
Gaussian and a non-Gaussian contribution:
\begin{subequations}
\label{deltaS4}
\begin{equation}
S_4 (t)= S_{\rm g}(t) +\delta S_4(t)
\label{deltaS4a}\,,
\end{equation}
where $S_{\rm g}$ is the Gaussian entropy as before, and where:
\begin{eqnarray}
\delta S_4(t) &=&
\frac{\eta_0 +\eta_0^* }{[2(a_{\mathrm{R}}\!-\!c)]^2}
\left(\frac{3n(1+n)}{1+2n}\ln\left(\frac{1+n}{n}\right)+\frac34\right)
+\frac34\frac{\eta_1+\eta_1^*}{[2(a_{\mathrm{R}}\!-\!c)]^2}
\left((1+2n)\ln\left(\frac{1+n}{n}\right)+1\right)
\nonumber\\
&& +\,\frac{\eta_2}{[2(a_{\mathrm{R}}\!-\!c)]^2}
\left( \frac{1+3n+3n^2}{1+2n}\ln\left(\frac{1+n}{n}\right)+\frac34\right)
\,.
\label{deltaS4b}
\end{eqnarray}
\end{subequations}
This is the main result of this section and intuitive in the
following sense: just as in the first non-Gaussian example above,
we see that positive kurtosis parameters
$\{\eta_{0R},\eta_{1R},\eta_2\}>0$, tend to make the effective
state's width $(a_{\mathrm{R}}-c)^{-1/2}$ larger, which increases
the area in phase space the state occupies, which in turn
increases the entropy. If however
$\{\eta_{0R},\eta_{1R},\eta_2\}<0$, the phase space area shrinks,
which in turn decreases the entropy. The statements above hold for
any statistical particle number $0\ll n<\infty$, with the
exception of $n\rightarrow 0$, where, just as in the case studied
above, a weak logarithmic divergence occurs when $\eta_{1R}\neq 0$
or $\eta_2\neq 0$. Notice finally that $\Im[\eta_0]$ and
$\Im[\eta_1]$ again do not participate in entropy generation, but
rather contribute to the squeezing of the state.
Kurtosis and skewness in quantum mechanics, studied in this
section, occur also in interacting quantum field theories which we
discuss next.
\section{Entropy in Scalar Field Theory}
\label{Entropy in scalar field theory}
The quantum mechanical expressions for the Gaussian and
non-Gaussian entropies that we have developed for pedagogical
reasons in sections \ref{Gaussian entropy from the Wigner
function}, \ref{Gaussian entropy from the replica trick}
and~\ref{Non-Gaussian entropy: Two examples} for a single particle
density matrix can be generalised to field theory. We firstly need
to consider correlators in quantum field theory however.
\subsection{Equal Time Correlators in Scalar Field Theory}
\label{Equal Time Correlators in Scalar Field Theory}
Let us now proceed analogous to equation~(\ref{density operator:
particle}) and write the density matrix operator for our system in
the field amplitude basis in Schr\"odinger's picture (see e.g.
\cite{Koksma:2007uq}):
\begin{equation}
\label{density operator: field} \hat \rho_{\mathrm{g}} = \int
{\cal D}\phi \int {\cal D}\phi^\prime
|\phi\rangle \rho_{\mathrm{g}}[\phi,\phi^\prime;t]\langle \phi^\prime| \,,
\end{equation}
where:
\begin{equation}
\label{density operator: field2}
\rho_{\mathrm{g}}[\phi,\phi^\prime;t] = {\cal N}
\exp\left(-\phi^T\!\cdot A\cdot\phi - {\phi^\prime}^T\!\cdot
B\!\cdot\!\phi^\prime + 2\phi^T\!\cdot C\!\cdot\!\phi^\prime
\right) \,,
\end{equation}
where ${\cal D}\phi = \prod_{\vec x\in V}\mathrm{d}\phi(\vec x\,)$
and $|\phi\rangle = \prod_{\vec x\in V}|\phi(\vec x\,)\rangle $. A
few words on the notation first. We shall consider $\phi=\phi(\vec
x)$ as a vector whose components are labelled by $\vec x$.
Moreover, $A(\vec x,\vec y,t)$ can be viewed as a matrix, such
that $A\cdot \phi$ is a vector again, where a $\cdot$ denotes
matrix multiplication which, for the case at hand, is nothing but
an integral over $D-1$ dimensional space. Hence, quantities like
$\phi^{\mathrm{T}}\cdot A \cdot \phi$ are scalars and involve two
integrals over space. Note that at this point we do not assume
that $A$ is homogeneous, i.e.: $A(\vec x,\vec y,t) \neq A(\vec x -
\vec y,t)$, but rather keep any possible off-diagonal terms for
generality.
Our density matrix (\ref{density operator: field2}) is hermitian,
such that we have $A^*(\vec x,\vec y,t)=B(\vec x,\vec y,t)$ and
$C^*(\vec x,\vec y,t)=C(\vec x,\vec y;t)$, where we also used
$A(\vec x,\vec y,t)=A(\vec y,\vec x,t)$ and $C(\vec x,\vec
y,t)=C(\vec y,\vec x;t)$. The normalisation ${\cal N}$ of
$\rho[\phi,\phi';t]$ can be determined from the standard
requirement ${\rm Tr}[\hat\rho]=1$:
\begin{equation}
{\cal N} = \left({\rm
det}\left[\frac{2(A_{\mathrm{R}}-C)}{\pi}\,\right]\right)^{1/2}
\,, \label{normrhofunc}
\end{equation}
where we note that the hermitian part $A_{\mathrm{h}}$ of $A$ is
real and symmetric such that $A_{\mathrm{h}}=A_{\mathrm{R}}$.
Just as in the quantum mechanical case, we need to calculate the
three non-trivial Gaussian correlators which completely
characterise the properties of our Gaussian state. In order to
calculate these correlators, we need to have an expression for the
statistical propagator as in equation (\ref{3 equal time
correlators}) which, given some initial density matrix
$\hat{\rho}(t_0)$, is in the Heisenberg picture defined by:
\begin{equation}
F(\vec x,t;\vec y,t')={\rm Tr}\left[\hat\rho(t_0)\hat\phi(\vec
x,t)\hat\phi(\vec y,t')\right] \label{statisticalpropagator} \,.
\end{equation}
Let us begin by calculating:
\begin{equation}
\langle \hat\phi(\vec x) \hat\phi(\vec y) \rangle = F(\vec x,\vec
y;t)={\rm Tr}\left[\hat\rho(t)\hat\phi(\vec x\,)\hat\phi(\vec
y\,)\right] \label{F:1} \,,
\end{equation}
where we have made use of the Heisenberg evolution equation for
operators. It is convenient to add a source current $j(\vec x,t)$
to the density matrix~(\ref{density operator: field}), such that
$\rho$ becomes:
\begin{equation}
\rho_{\mathrm{g}}^{j}[\phi,\phi^\prime;t] = {\cal N}
\exp\left(-\phi^T\!\cdot A\cdot\phi - {\phi^\prime}^T\!\cdot
A^{\dag}\!\cdot\!\phi^\prime + 2\phi^T\!\cdot
C\!\cdot\!\phi^\prime + j^T\!\cdot \phi + {j'}^T\!\cdot
\phi^\prime \right) \,, \label{density operator: field:j}
\end{equation}
in terms of which equation (\ref{F:1}) can be rewritten as:
\begin{subequations}
\label{correlatorsQFT}
\begin{eqnarray}
F(\vec x,\vec y;t) &=&\frac{\delta}{\delta j(\vec x\,)}
\frac{\delta}{\delta j(\vec y\,)} \int {\cal D}\phi
\rho^j_{\mathrm{g}}[\phi,\phi;t]\Big|_{j=j^\prime=0}
\nonumber\\
&=&\frac{\delta}{\delta j(\vec x\,)}\frac{\delta}{\delta j(\vec
y\,)}\int {\cal D}\tilde\phi \, {\cal N} \left.
\exp\left(-2\tilde\phi^T\!\cdot(A_{\mathrm{R}}-C)\cdot\tilde
\phi\right)\exp\left(\frac18(j+j^\prime)^T\!\cdot(A_{\mathrm{R}}-C)^{-1}
\cdot(j+j^\prime)\right) \right|_{j=j^\prime=0}
\nonumber\\
&=& \frac14(A_{\mathrm{R}}-C)^{-1}(\vec x,\vec y;t)\,.
\label{correlatorsQFTa}
\end{eqnarray}
We also need the other correlators:
\begin{eqnarray}
\frac{1}{2} \langle \{ \hat{\phi}(\vec{x}), \hat{\pi}(\vec{y}) \}
\rangle &=& \partial_{t'} F(\vec x,t;\vec
y,t^\prime)|_{t=t^\prime} = \frac12{\rm Tr}[\hat
\rho_{\mathrm{g}}(t) \{\hat\phi(\vec x\,),\hat\pi(\vec y\,)\}]
=-\frac{\hbar}{2}(A_{\mathrm{R}}-C)^{-1}\cdot A_{\mathrm{I}}(\vec
x,\vec y;t)
\label{correlatorsQFTb} \\
\frac{1}{2} \langle \{ \hat{\pi}(\vec{x}), \hat{\pi}(\vec{y}) \}
\rangle &=& \partial_t\partial_{t^\prime}F(\vec x,t;\vec
y,t^\prime)|_{t=t^\prime} = \frac12{\rm Tr}[\hat\rho_{\mathrm{g}}
(t) \{\hat\pi(\vec x\,),\hat\pi(\vec y\,)\}]
\label{correlatorsQFTc}\\
&=& \hbar^2 \! \left[ \frac{1}{2} A^{\dag} \! \cdot\!
(A_{\mathrm{R}} \! -C)^{-1} \! \cdot \!A \!+\! A \cdot\!
(A_{\mathrm{R}}\!-C)^{-1} \! \cdot A^{\dag}\! -C \!\cdot\!
(A_{\mathrm{R}}\!-C)^{-1}\! \cdot C \!\right](\vec x,\vec y;t)
\nonumber\,.
\end{eqnarray}
\end{subequations}
We have moreover made use of $\langle\phi'|\hat{\pi}|\phi\rangle =
-\imath \hbar \frac{\delta}{\delta\phi'} \delta[\phi'-\phi]$. As a
check one can verify that:
\begin{equation}
{\rm Tr}(\hat \rho_{\mathrm{g}}(t)[\hat\phi(\vec x\,),\hat\pi(\vec
y\,)]) = \imath\hbar\delta^{\scriptscriptstyle{D}-1}(\vec x-\vec
y\,) \,. \label{commutator}
\end{equation}
Combining the equations above we find:
\begin{eqnarray}
\Delta^2(\vec x,\vec y;t) &=& \frac{4}{\hbar^2} \left[ \left
\langle\hat\phi\hat\phi\right\rangle \cdot \left
\langle\hat\pi\hat\pi \right \rangle- \frac14 \left\langle
\{\hat\phi,\hat\pi\}\right\rangle \cdot\left\langle
\{\hat\phi,\hat\pi\}\right\rangle \right](\vec x,\vec y;t)
\nonumber\\
&=& (A_{\mathrm{R}}-C)^{-1}\cdot(A_{\mathrm{R}}+C)(\vec x,\vec y;t)
\,. \label{Delta function:phi}
\end{eqnarray}
This is the desired field theoretic generalisation of the Gaussian
invariant $\Delta^2$ in equation (\ref{Delta function:3}). Notice
that the result above applies for general non-diagonal Gaussian
density matrices. This is an important quantity because, just as
in the quantum mechanical case, $\Delta$ will be the conserved
quantity under any quadratic Hamiltonian evolution. Since the von
Neumann entropy is also conserved in this case, it is natural to
expect that $S_{\mathrm{vN}}=S_{\mathrm{vN}}[\Delta]$.
Of course, if we are interested in problems in which the
hamiltonian density is only time dependent, one can make use of
spatial translation invariance of the correlators, such that the
equal time statistical correlator is homogeneous: $F(\vec x, \vec
y;t) \rightarrow \tilde F(\vec x-\vec y,t)$. In this case it is
beneficial to Fourier transform according to:
\begin{equation}
\label{Fouriertransform} \tilde F (\vec k,t) = \int
\mathrm{d}^{\scriptscriptstyle{D}-1}(\vec x - \vec y)
\tilde F(\vec x-\vec y,t) \mathrm{e}^{-\imath\vec
k\cdot(\vec x - \vec y)}\,,
\end{equation}
such that equation (\ref{Delta function:phi}) becomes local in
momentum space and reduces to the result known in the literature:
\begin{eqnarray}
\tilde{\Delta}^2(\vec k;t) &=& \frac{4}{\hbar^2} \left.
\left[\tilde F(\vec k,t,t)
\partial_t\partial_{t^\prime}\tilde F(\vec k,t,t^\prime) -
\left(\partial_{t'} \tilde F(\vec k,t,t^\prime) \right)^2 \right]
\right|_{t=t^\prime}
\nonumber\\
&=&\frac{(\tilde A_{\mathrm{R}}+ \tilde C)(\vec k,t)}{( \tilde
A_{\mathrm{R}}- \tilde C)(\vec k,t)} \,. \label{Delta
function:phi:momentum}
\end{eqnarray}
This representation is particularly useful in problems with
spatial translational symmetry, such as
cosmology~\cite{Brandenberger:1992jh,Campo:2008ij}.
\subsection{Entropy of a Gaussian state in Scalar Field Theory}
\label{Entropy of a Gaussian state in scalar field theory}
Let us now discuss the von Neumann entropy~(\ref{entropy:vN}) of
the Gaussian density matrix~(\ref{density operator: field}) by
using the replica method~(\ref{replica trick}). One can proceed
analogous to section~\ref{Gaussian entropy from the replica
trick}. Some subtleties arise however as we deal with a system
with infinite degrees of freedom. For this reason, we nevertheless
include an outline of the proof in appendix~\ref{Appendix B
Entropy for Gaussian Field Theory}. The entropy for a quantum
system that can be described by a Gaussian density matrix is given
by:
\begin{equation}
S_{\rm vN} = {\rm
Tr}\left[\frac{\Delta+\mathbb{I}}{2}\cdot\ln\left(\frac{\Delta+\mathbb{I}}{2}\right)
- \frac{\Delta-\mathbb{I}}{2}\cdot
\ln\left(\frac{\Delta-\mathbb{I}}{2}\right) \right] \,.
\label{entropy:vN:Delta}
\end{equation}
We denote the identity matrix by
$\mathbb{I}=\delta^{\scriptscriptstyle{D}-1}(\vec x -\vec y)$. As
in the quantum mechanical case, we can define the generalised
statistical particle number density correlator $n=n(\vec x,\vec
y,t)$ as:
\begin{equation}
n = \frac{\Delta -\mathbb{I}}{2} \label{n:matrix}\,,
\end{equation}
in terms of which the entropy~(\ref{entropy:vN:Delta}) reads:
\begin{equation}
S_{\rm vN} = {\rm Tr}\left[(n+\mathbb{I})\cdot\ln
\left(n+\mathbb{I} \right) - n\cdot\ln \left(n \right) \right] \,.
\label{entropy:vN:n}
\end{equation}
In the limit when $\|n \| \gg 1$, in the sense that for the
diagonalised matrix its diagonal entries of interest are large, we
can expand the logarithms in~(\ref{entropy:vN:n}) to get for the
entropy, $S_{\rm vN} \approx {\rm Tr}[\ln(n)+\mathbb{I} +{\cal
O}(n^{-1})]$, which nearly coincides with the Wigner entropy
$S_{\cal W}$ in field theory, {\it cf.} equation~(\ref{Wigner
function:entropy:2}). Equations~(\ref{entropy:vN:n})
and~(\ref{entropy:vN:Delta}) represent the von Neumann entropy of
a general Gaussian state in scalar field theory and are the main
result of this section.
In the homogeneous limit, we can again Fourier transform and
equation (\ref{entropy:vN:Delta}) reduces to:
\begin{equation}
S_{\rm vN} = V \int \frac{\mathrm{d}^{3}\mathbf{k}}{(2\pi)^3}
\left[\frac{\Delta (\mathbf{k},t)+1}{2}
\ln\left(\frac{\Delta(\mathbf{k},t)+1}{2}\right) -
\frac{\Delta(\mathbf{k},t)-1}{2}
\ln\left(\frac{\Delta(\mathbf{k},t)-1}{2}\right) \right] \,,
\label{entropy:vN:Delta_FourierTransform}
\end{equation}
where the volume factor arises because of the trace. The limit
$\Delta(\mathbf{k},t) \gg 1$ basically agrees with
\cite{Brandenberger:1992jh}.
\subsection{Entropy of a Non-Gaussian state in Scalar Field Theory}
\label{Entropy of simple non-Gaussian states in scalar field
theory}
Analogous to the one particle non-Gaussian entropy discussed in
section~\ref{Non-Gaussian entropy: Two examples}, let us
generalise that result to the field theoretical case. The
non-Gaussian density matrix $\hat \rho_2$ in
equation~(\ref{density operator: field}) generalises to:
\begin{subequations}
\label{density operator: field:nonG}
\begin{equation}
\hat \rho_2 = \int {\cal D}\phi \int {\cal D}\phi^\prime
|\phi\rangle \rho_2[\phi,\phi^\prime;t]\langle \phi^\prime| \,,
\label{density operator: field:nonGa}
\end{equation}
where:
\begin{eqnarray}
\rho_2[\phi,\phi^\prime;t] &=& {\cal N}_2\left[1+\zeta_0\cdot\phi
+ \zeta_0^\dagger\cdot\phi^\prime +\phi^T\cdot\zeta_1\cdot\phi +
{\phi^\prime}^T\cdot\zeta_1^\dagger\cdot\phi^\prime
+\phi^T\cdot\zeta_2\cdot{\phi^\prime} +
{\phi^\prime}^T\cdot\zeta_2^\dagger\cdot\phi
\right]\rho_0[\phi,\phi^\prime;t] \label{density operator:
field:nonGb}
\\
\rho_0[\phi,\phi^\prime;t] &=& \exp\left[-\phi^T\!\cdot A\cdot\phi
- {\phi^\prime}^T\!\cdot A^{\dag}\!\cdot\!\phi^\prime
+ \phi^T\!\cdot C\!\cdot\!\phi^\prime
+ {\phi^\prime}^T\!\cdot C^T\cdot\!\phi
\right] \,,
\label{density operator: field:nonGc}
\end{eqnarray}
\end{subequations}
We could of course perform a similar shift in the exponent as we
have done before in equation (\ref{a+c:rescaling}) in which case
the resulting equation for the entropy would formally be exact,
i.e: we do not assume yet that the non-Gaussian contributions are
small. If we want to expand around the Gaussian result however, we
can only perform the integral if we assume that all correlators,
including the non-Gaussian ones, are homogeneous, i.e.: they are
only a function of the difference of their coordinates $A(\vec
x,\vec y,t)=A(\vec x-\vec y,t)$. The Fourier transform of equation
(\ref{density operator: field:nonG}) is:
\begin{equation}
\rho_2 [\phi,\phi^\prime;t] = \prod_{\mathbf{k}} {\cal
N}_{2,\mathbf{k}} \exp\left[- \bar
A(\mathbf{k},t)|\phi(\mathbf{k})|^{2}- \bar
A^{\ast}(\mathbf{k},t)|\phi'(\mathbf{k})|^{2} + 2 \bar
C(\mathbf{k},t)\phi^{\ast}(\mathbf{k}) \phi'(\mathbf{k}) \right] +
{\cal O}(\zeta_i^2) \,, \label{density operator: field:nonG2}
\end{equation}
where we have set $\zeta_0=0$ as before, as it will not induce an
entropy shift. Also, the product over the momenta is only over
half of the Fourier space. Moreover, we have absorbed the small
non-Gaussian contributions in the functions $\bar A$ and $\bar C$
as in the quantum mechanical case:
\begin{subequations}
\label{a+c:rescalingQFT}
\begin{eqnarray}
A(\mathbf{k},t) &\rightarrow & \bar A(\mathbf{k},t) = A(\mathbf{k},t) - \zeta_1 (\mathbf{k},t) \label{a+c:rescalingQFTa}\\
C(\mathbf{k},t) &\rightarrow & \bar C(\mathbf{k},t) =
C(\mathbf{k},t) + \zeta_2 (\mathbf{k},t)
\label{a+c:rescalingQFTb}\,,
\end{eqnarray}
\end{subequations}
Now we can read off the result for the entropy in Fourier space in
equation (\ref{entropy:vN:Delta_FourierTransform}) as:
\begin{equation}
S_{\rm vN} = V \int \frac{\mathrm{d}^{3}\mathbf{k}}{(2\pi)^3}
\left[\frac{\bar \Delta (\mathbf{k},t)+1}{2} \ln\left(\frac{\bar
\Delta(\mathbf{k},t)+1}{2}\right) - \frac{\bar
\Delta(\mathbf{k},t)-1}{2} \ln\left(\frac{\bar
\Delta(\mathbf{k},t)-1}{2}\right) \right] \,,
\label{entropy:vN:Delta_FourierTransformNGcase}
\end{equation}
Assuming that $\zeta_i(\mathbf{k},t) \ll 1$ we can derive the
change in entropy to linear order in $\zeta_i$. The result is:
\begin{subequations}
\label{s2:finalQFT}
\begin{equation}
S_2(t)=S_{\mathrm{g}}(t) + \delta S_{2}(t) \label{s2:finalQFTa}
\,,
\end{equation}
where $S_{\mathrm{g}}$ is the Gaussian contribution and $\delta
S_{2}$ is an entropy shift given by:
\begin{eqnarray}
\delta S_2 &=& V \int \frac{\mathrm{d}^{3}\mathbf{k}}{(2\pi)^3}
\Bigg[\left(\frac{\zeta_{1R}(\mathbf{k},t)}{A_{\mathrm{R}}(\mathbf{k},t)-C(\mathbf{k},t)}+
\frac{1+2n(\mathbf{k},t)+2n^2(\mathbf{k},t)}{2n(\mathbf{k},t)(1+n(\mathbf{k},t))}\frac{\zeta_2(\mathbf{k},t)}{A_{\mathrm{R}}(\mathbf{k},t)-C(\mathbf{k},t)}\right)
\nonumber \\
&& \qquad\qquad\qquad \times
\frac{n(\mathbf{k},t)(1+n(\mathbf{k},t))}{1+2n(\mathbf{k},t)}
\ln\left(\frac{1+n(\mathbf{k},t)}{n(\mathbf{k},t)}\right)\Bigg]
\label{s2:finalbQFT}\,,
\end{eqnarray}
\end{subequations}
where, of course, $n(\mathbf{k},t) = (\Delta(\mathbf{k},t)-1)/2$.
This result is the field theoretic generalisation of the quantum
mechanical entropy~(\ref{s2:final}). It can be applied to mildly
non-Gaussian states in field theory, which, apart from being
mixed, also contain small one particle contributions.
The field theoretical generalisation of the second example
presented in section \ref{Kurtosis} is hard to solve for, even in
the homogeneous case, as it is non-local in Fourier space.
\subsection{Example: A Scalar Field with a Changing Mass}
\label{Example: A Scalar Field with a Changing Mass}
As a simple illustration of the ideas presented above, let us
investigate the effect of a changing mass on the Gaussian entropy
for a scalar field. This is maybe not a very exciting example, as
no entropy is generated of course in free theories, it
nevertheless provides an intuitive way of how the Wigner function
can be used. Let us consider the action of a free scalar field:
\begin{equation}\label{action1}
S[\phi] = \int \mathrm{d}^{4}\!x \left\{-\frac{1}{2}
\partial_\mu\phi(x)
\partial_\nu \phi(x) \eta^{\mu\nu} - \frac{1}{2} m^{2}_{\phi}(t)
\phi^{2}(x) \right\}\,,
\end{equation}
where as usual $\eta_{\mu\nu} = {\rm diag}(-1,1,1,1)$ is the
Minkowski metric, and where we consider the following behaviour of
the mass $m_{\phi}(t)$ of the scalar field, mediated for example
by some other Higgs-like scalar field:
\begin{equation}\label{masstanh}
m^{2}_{\phi}(t) = \left(A + B \tanh(\rho t)\right)\,.
\end{equation}
The equation of motion following from (\ref{action1}) reads:
\begin{equation} \label{eom1}
\left(\partial_{t}^{2}-\partial_{i}^{2}+m^{2}_{\phi}(t) \right)
\phi(x) =0\,,
\end{equation}
Let us quantise our fields in $D$-dimensions by making use of
creation and annihilation operators:
\begin{equation} \label{quantisationphi}
\hat{\phi}(x) = \int
\frac{\mathrm{d}^{\scriptscriptstyle{D}-1}\vec{k}}{(2\pi)^{\scriptscriptstyle{D}-1}}
\left( \hat{a}_{\vec{k}}\,\phi_{k}(t)e^{i\vec{k} \cdot \vec{x}} +
\hat{a}_{\vec{k}}^{\dag}\,\phi_{k}^{\ast}(t)e^{-i\vec{k} \cdot
\vec{x}} \right)\,.
\end{equation}
The annihilation operator acts on the vacuum as usual
$\hat{a}_{\vec{k}}|0\rangle = 0$. We impose the following
commutation relations: $[ \hat{a}_{\vec{k}},
\hat{a}_{\vec{k'}}^{\dag}] =(2\pi)^{\scriptscriptstyle{D}-1}
\delta^{\scriptscriptstyle{D}-1}(\vec{k}-\vec{k'})$. The mode
functions $\phi_{k}(t)$ of $\phi(x)$, defined by relation
(\ref{quantisationphi}) thus obey:
\begin{equation}\label{eom2}
\left(\partial_{t}^{2} + \omega^{2}(t) \right) \phi_{k}(t) = 0 \,,
\end{equation}
where $k=\|\vec{k}\|$ and $\omega^{2}(t)=k^{2} + m^{2}_{\phi}(t)$.
Using equation (\ref{F:1}) with
$\hat{\rho}(t_{0})=|0\rangle\langle0|$, we see that the mode
functions determine the statistical propagator completely:
\begin{equation}
F_{\phi}(k,t,t') = \frac{1}{2} \left\{
\phi_{k}(t')\phi_{k}^{\ast}(t) +
\phi_{k}(t)\phi_{k}^{\ast}(t')\right\}
\label{statisticalpropagatorF} \,.
\end{equation}
The solution of (\ref{eom2}) which behaves as a positive frequency
mode in the asymptotic past, i.e.: $\lim_{t\rightarrow - \infty}
\phi_{k}^{\mathrm{in}}(t) = \exp\left[- \imath
\omega_{\mathrm{in}} t \right]/\sqrt{2\omega_{\mathrm{in}}}$, can
be expressed in terms of Gauss' hypergeometric function
${}_{2}F_{1}$ (see \cite{Bernard:1977pq, Birrell:1982ix}):
\begin{equation} \label{modesolution1}
\phi_{k}^{\mathrm{in}}(t) = \frac{1}{\sqrt{2\omega_{\mathrm{in}}}}
\exp \left[- \imath \omega_{+} t - \imath
\frac{\omega_{-}}{\rho}\log\{2\cosh(\rho t )\}\right]
\phantom{1}_{2}F_{1}\left( 1+ \imath \frac{\omega_{-}}{\rho},
\imath \frac{\omega_{-}}{\rho}; 1- \imath
\frac{\omega_{\mathrm{in}}}{\rho};\frac{1}{2}\{1+\tanh(\rho
t)\}\right) \,,
\end{equation}
where we defined $\omega_{\mathrm{in}} = ( k^{2}+A-B
)^{\frac{1}{2}}$, $\omega_{\mathrm{out}} = (
k^{2}+A+B)^{\frac{1}{2}}$ and $\omega_{\pm}= (
\omega_{\mathrm{out}} \pm \omega_{\mathrm{in}})/2$. Having the
mode functions at our disposal, we can find the rather cumbersome
expressions for the exact statistical propagator. The statistical
propagator in turn fixes the phase space area through equation
(\ref{Delta function:phi:momentum}), yielding $\Delta_{k}(t) = 1$
such that:
\begin{equation}\label{solS2}
S_{k}(t) = 0 \,.
\end{equation}
\begin{figure}[t!]
\begin{minipage}[t]{.43\textwidth}
\begin{center}
\includegraphics[width=\textwidth]{Wigner1.eps}
{\em \caption{Squeezing of the Gaussian phase space in the Wigner representation
due to a changing mass. We
used $k/\rho= 0.2$ and $m/\rho$ changes from 0 to 4 where the mass changes most rapidly at $t\rho=5$. For early
times $t\rho=0$, the Gaussian phase space is a perfect circle (blue).
Already for $t\rho=2$ a little squeezing is visible (red) and at late
times $t\rho=9$ this is manifest (black). The coordinates $p_{r}$ and $q_{r}$
have been rescaled for dimensional reasons:
$p_{r}=p\sqrt{\omega(t)}$ and $q_{r}=q/\sqrt{\omega(t)}$.
\label{fig:Wigner1}}}
\end{center}
\end{minipage}
\hfill
\begin{minipage}[t]{.43\textwidth}
\begin{center}
\includegraphics[width=\textwidth]{Wigner2.eps}
{\em \caption{Rotation of the Gaussian phase space in the Wigner representation
due to a mass that has changed. We used the same parameters as
in figure \ref{fig:Wigner1}. The squeezed phase space ellipses
are plotted at times $t\rho=9.5$ (green), $t\rho=9.6$ (blue),
$t\rho=9.7$ (red) and $t\rho=9.8$ (black). The rotation is clearly visible and its direction is indicated by the arrow. This can be
understood from realising that an arbitrary squeezed state can
be interpreted as a superposition of ordinary coherent states
and the knowledge that the latter show the same rotating
behaviour. \label{fig:Wigner2}}}
\end{center}
\end{minipage}
\end{figure}
We thus conclude that a changing mass does not change the entropy
for a free scalar field. Let us now examine the same process in
Wigner space. Of course, we will reach a similar conclusion
(\ref{solS2}) but Wigner space is much more suited to visualise
the process neatly. The statistical propagator is also the
essential building block for the Wigner function which can be
appreciated from generalising equations (\ref{aI:aR:c}) and
(\ref{Wigner function:Gauss2}) to:
\begin{subequations}
\label{Wignercoefficients}
\begin{eqnarray}
\alpha_{\mathrm{w}}(k,t) &=& \frac{1}{2 F_{\phi}(k,t,t)} \label{Wignercoefficientsa}\\
\beta_{\mathrm{w}}(k,t) &=& \frac{2
F_{\phi}(k,t,t)}{\Delta_{k}^{2}(t)}
\label{Wignercoefficientsb}\\
p_{\mathrm{c}} &=& - \frac{\partial_{t}
F_{\phi}(k,t,t')|_{t=t'}}{F_{\phi}(k,t,t)}
\label{Wignercoefficientsc} \\
\mathcal{M}_{k}(t) &=& \frac{2}{\Delta_{k}(t)}
\label{Wignercoefficientsd} \,.
\end{eqnarray}
\end{subequations}
We can thus plot cross-sections of the 2-dimensional Gaussian
phase space distribution in the Wigner representation. In figure
\ref{fig:Wigner1} we depict the squeezing of the Gaussian phase
space due to the changing mass. At early times, the state is still
in an unsqueezed vacuum. At $t\rho=2$ some squeezing is visible,
whereas at late times $t\rho=9$ this is manifest. Despite the
effect of the changing mass on the accessible phase space in
Wigner space, the area of the ellipse remains constant throughout
the whole process as expected from equation (\ref{solS2}). At late
times, when the mass has settled to its final constant value
$m/\rho=4$, the squeezed ellipse rotates in Wigner space, which we
depict in figure \ref{fig:Wigner2}, but it is not squeezed any
further. The rotation can be anticipated by the intuitive notion
that an arbitrary squeezed state can be thought of as a
superposition of coherent states (just imagine that the ellipse is
replaced by a number of circles displaced from the origin).
Coherent states that are displaced from the origin rotate in time.
\section{Conclusion}
\label{Conclusion}
We have formally developed a novel approach to decoherence in
quantum field theory: neglecting observationally inaccessible
correlators will give rise to an increase in the entropy of the
system. This is inspired by realising that correlators are
measured in quantum field theories and that higher order n-point
functions are usually perturbatively suppressed. An important
advantage of this approach is that the procedure of
renormalisation can systematically be implemented in this
framework.
We have shown how knowledge about the correlators of a system
affects the notion of the entropy associated to that state in two
cases: we firstly calculated the entropy of a general Gaussian
state. This entropy can be expressed purely in terms of the three
equal time correlators characterising the Gaussian state.
Moreover, all three correlators can be obtained from the
statistical propagator which opens the possibility to study
quantum corrections on the entropy in an interacting quantum field
theory in a systematic manner. Secondly, we calculated the entropy
for two specific types of non-Gaussian states. Firstly, we assumed
that the state of the system can be described by an admixture of
the ground state and a small contribution of the first excited
state. This yields a small correction to the Gaussian entropy.
Secondly, in the quantum mechanical case, we calculated an
expression for the entropy in case our observer could probe a
specific type of kurtosis of the ground state. This also changes
the entropy that the observes associates to the state.
We also outlined the use and limitations of the Wigner function,
the Wigner transform of the density matrix of a system. In
particular, we have shown that $\Delta$, the fundamental quantity
constructed from various correlators that fixes the entropy,
indeed coincides with the phase space area in Wigner space.
This is a rather phenomenological discussion, connecting the
notion of entropy, correlators and phase space area in quantum
field theory. Although we have outlined various mechanisms how
entropy could be generated, we have in the present paper not
applied our ideas to concrete systems. We refer the reader to e.g.
\cite{Giraud:2009tn, Koksma:2009wa, Koksma:2010} for specific
examples of how entropy is generated in interacting quantum field
theories in an out-of-equilibrium setting.
\
\noindent \textbf{Acknowledgements}
\noindent JFK and TP thank Theo Ruijgrok and gratefully
acknowledge the financial support from FOM grant 07PR2522 and by
Utrecht University. The authors also gratefully acknowledge the
hospitality of the Nordic Institute for Theoretical Physics
(NORDITA) during their stay at the ``Electroweak Phase
Transition'' workshop in June, 2009.
|
\section{Lattice Paths with Constraints}
Lattice paths arise in several areas in probability and combinatorics, either in their own interest
(as realizations of random walks, or because of their interesting combinatorial properties:
see \cite{Ban} for the latter) or because of fruitful bijections with
various families of trees, tilings, words.
The problem we discuss here is to efficiently sample uniform (or \emph{almost} uniform)
paths in a family of paths with constraints.
There are several reasons for which one may want to generate uniform samples of lattice paths:
to make and try conjectures on the behaviour of a large "typical" path, test
algorithms running on paths (or words, trees,...).
In view of random sampling, it is often very efficient to make use of the combinatorial structure of the
family of paths under study. In some cases, this yields linear-time (in the length of the path)
\emph{ad-hoc} algorithms \cite{MBM,Duc}. However, the nature of the constraints makes sometimes impossible
such an approach, and there is a need for robust algorithms that work in lack of
combinatorial knowledge.
Luby,Randall and Sinclair \cite{LRS} design a Markov chain that generate
sets of non-intersecting lattice paths. This was motivated by a classical (and simple, see illustrations in \cite{Des,Wilson}) correspondence between dimer configurations on an hexagon, rhombae tilings of this hexagon and families of non-intersecting lattice paths.
As the first step for the analysis of this chain, Wilson \cite{Wilson} introduces a peak/valley
Markov chain (see details below) over some simple lattice paths and obtain sharp bounds for its mixing
time.
We present in this paper a variant of this Markov chain,
which is valid for various constraints and whose analysis is simple.
It generates an "almost" uniform path of length $n$ in $n^{3+\varepsilon}$ steps, this bound makes use of a certain
\emph{contraction property} of the chain.
Appart from the algorithmic aspect, the peak/valley process seems to have a physical relevancy as a simplified model for the evolution of \emph{quasicrystals} (see a discussion on a related process in the introduction of \cite{Des}). In particular, the mixing time of this Markov seems to have some importance.
\subsection*{Notations}
\begin{figure}[h]
\begin{center}
\includegraphics[width=40mm]{ExempleChemin.eps}
\caption{The lattice path $S=(1,2,0,1,2,3,1)$ associated with the word $(1,1,-2,1,1,1,-2)$.}
\end{center}
\end{figure}
We fix three integers $n,a,b>0$, and consider the paths of length $n$, with steps $+a/-b$,
that is, the words of $n$ letters taken in the alphabet $\set{a,-b}$.
Such a word $s=(s_1,s_2,\dots,s_n)$ is identified to the path
$S=(S_1,\dots,S_n):=(s_1,s_1+s_2,\dots,s_1+s_2+\dots +s_n)$.
To illustrate the methods and the results, we focus on some particular sub-families
$\SsChemins{n}\subset \set{a,-b}^n$:
\begin{enumerate}
\item Discrete \emph{meanders}, denoted by $\Meandres{n}$, which are simply the non-negative paths: $S\in\Meandres{n}$ if for
any $i\leq n$ we have $S_i\geq 0$. This example is mainly illustrative because the combinatorial properties of meanders
make it possible to perform exact sampling very efficiently (an algorithm running in $\mathcal{O}(n^{1+\varepsilon})$
steps is given in \cite{MBM}, an order that we cannot get in the present paper).
\item Paths with \emph{walls}. A path with a wall of height
$h$ between $r$ and $s$ is a path such that
$S_i\geq h$ for any $r\leq i\leq s$ (see Fig. \ref{Fig:CheminMur} for an example).
These are denoted by $\Walls{n}=\Walls{n}(h,r,s)$, they
appear in statistical mechanics as toy models for the analysis of random interfaces
and polymers (see examples in \cite{Walls}).
\item \emph{Excursions}, denoted by $\Exc{n}$, which are non-negative paths such that $S_n=0$.
In the case $a=b=1$, these correspond to well-parenthesed words and are usually called Dyck words.
In the general case, Duchon \cite{Duc} proposes a rejection algorithm which generates excursions
in linear time.
\item \emph{Culminating paths} of size $n$, denoted further by $\Culmis{n}$, which are non-negative paths whose maximum is attained at the last step: for any $i$ we have $0\leq S_i\leq S_n$. They have been introduced in \cite{MBM},
motivated in particular by the analysis of some algorithms in bioinformatics.
\end{enumerate}
\begin{figure}[h]
\begin{center}
\includegraphics[width=65mm]{CheminMur.eps}
\caption{A path of steps $+1/-2$, with a wall of height $h=6$ between $i=10$ and $j=15$.}
\label{Fig:CheminMur}
\end{center}
\end{figure}
\section{Sampling with Markov chains}\label{Sec:Sampling}
We will consider Markov chains in a family $\SsChemins{n}$, where all the probability transitions are symmetric. For a modern introduction to Markov chains, we refer to \cite{Hagg}.
Hence we are given a transition matrix $(p_{i,j})$ of size $|\SsChemins{n}|\times|\SsChemins{n}|$ with
\begin{align*}
p_{i,j} &=p_{j,i} \mbox{ whenever }i\neq j,\\
p_{i,i} &= 1-\sum_{j\neq i}p_{i,j}.
\end{align*}
\begin{lem}\label{Lem:Unif}
If such a Markov chain is irreducible, then it admits as unique stationary distribution
the uniform distribution $\pi=\pi(\SsChemins{n})$ on $\SsChemins{n}$.
\end{lem}
\begin{proof}
The equality $\pi(i) p_{i,j}= \pi(j) p_{j,i}$ holds for any two vertices $i,j$. This shows that
the probability distribution $\pi$ is reversible for $(p_{i,j})$, and hence stationary. It is unique
if the chain is irreducible.
\end{proof}
This lemma already provides us with a scheme for sampling an almost uniform path in $\SsChemins{n}$, without knowing much about $\SsChemins{n}$.
To do so, we define a ``flip'' operator on paths, this is an operator
$$
\begin{array}{r c c c}
\phi: & \SsChemins{n}\times \set{1,\dots,n}\times \set{\downarrow,\uparrow}\times \set{+,-} &\to &\SsChemins{n}\\
& (\mathbf{S},i,\varepsilon,\delta) &\mapsto & \phi(\mathbf{S},i,\varepsilon,\delta).
\end{array}
$$
When $i\in\set{1,2,\dots,n-1}$ the path $\phi(\mathbf{S},i,\uparrow,\delta)$ is defined as follows :
if $(s_i,s_{i+1})=(-b,a)=$ \includegraphics[width=7mm]{downup.eps}
then these two steps are changed into $(a,-b)=$ \includegraphics[width=7mm]{updown.eps}. The $n-2$ other steps remain unchanged.
If $(s_i,s_{i+1})\neq (-b,a)$ then $\Flip{S}{i}{\uparrow}{\delta}=\mathbf{S}$. Note that in the case
$i\in\set{1,2,\dots,n-1}$ the value of $\phi$ does not depend on $\delta$.
For the case $i=n$, if $\delta=+$, we define $\Flip{S}{n}{\varepsilon}{\delta}$ as before as if there would be a $+a$ as the
end if the path. For instance, in the case where $S_n=-b$, the path $\Flip{S}{n}{\uparrow}{+}$, the $n$-th step is turned into $a$.
The path $\Flip{S}{i}{\downarrow}{\delta}$ is defined equally: if $i<n$ and $(s_i,s_{i+1})=$ \includegraphics[width=7mm]{updown.eps},
it turns into \includegraphics[width=7mm]{downup.eps}. When $\delta=-$, one flips as if there would be a
$-b$ at the end of the path.
For culminating paths, we have to take another definition
of $\Flip{S}{n}{\uparrow}{\delta},\Flip{S}{n}{\downarrow}{\delta}$, see Section \ref{Sec:Analysis}.
We are also given a probability distribution $\mathbf{p}=(p_i)_{1\leq i\leq n}$, and we assume that $p_i>0$ for each $i$.
We will consider a particular sequence $\mathbf{p}$ later on, but at this point we can take
the uniform distribution in $\set{1,\dots,n}$. We describe the algorithm below in Algorithm \ref{Algo:CM}.
\begin{algorithm}
\caption{Approximate sampling of a path in $\SsChemins{n}$}
\label{Algo:CM}
\begin{algorithmic}
\STATE initialize $\mathbf{S}=(+a,+a,+a,\dots,+a)$
\STATE $I_{1},I_{2},\dots\leftarrow$ i.i.d. r.v. with law $\mathbf{p}$
\STATE $\varepsilon_{1},\varepsilon_{2},\dots\leftarrow$ i.i.d. uniform r.v. in $\set{\uparrow,\downarrow}$
\STATE $\delta_{1},\delta_{2},\dots\leftarrow$ i.i.d. uniform r.v. in $\set{+,-}$
\FOR{$t=1$ to $T$}
\IF{$\Flip{S}{I_t}{\varepsilon_t}{\delta_t}$ is in $\SsChemins{n}$} \STATE $\mathbf{S}\leftarrow
\Flip{S}{I_t}{\varepsilon_t}{\delta_t}$
\ENDIF
\ENDFOR
\end{algorithmic}
\end{algorithm}
In words, this algorithm performs the Markov chain in $\SsChemins{n}$ with transition matrix
$P=\left(P_{\mathbf{R},\mathbf{S}}\right)_{\mathbf{R},\mathbf{S}\in\SsChemins{n}}$ defined as follows:
$$
\begin{cases}
P_{\mathbf{R},\mathbf{S}}&=p_i/2, \mbox{ if } \mathbf{S}\neq\mathbf{R}\text{ and } \mathbf{S}=\Flip{R}{i}{\varepsilon}{\delta} \mbox{ for some }i,\varepsilon, \delta\\
P_{\mathbf{R},\mathbf{S}}&=0\text{ otherwise,}\\
P_{\mathbf{R},\mathbf{R}}&=1-\sum_{\mathbf{S}\neq \mathbf{R}} P_{\mathbf{R},\mathbf{S}}.\\
\end{cases}
$$
\begin{prop}
Denote by $S(t)$ the random path obtained after the $t$-th run of the loop in
Algorithm \ref{Algo:CM}. When $t\to\infty$, the
sequence $S(t)$ converges in law to the uniform distribution in $\SsChemins{n}$.
Moreover, the execution of Algorithm \ref{Algo:CM} until time $T$ is linear in $T$.
\end{prop}
\begin{proof}
For the first claim, we have to check that the chain is aperiodic and irreducible.
Aperiodicity comes from the (many) loops. Irreducibility will follow from
Lemma \ref{Lem:Geodesique}.
For the second claim, notice that the time needed for the test "$\Flip{S}{I_t}{\varepsilon_t}$ is in $\SsChemins{n}$" can be considered as constant, since for the families $\Meandres{n}$ and $\Exc{n}$ we only have to
compare $0,S_i$ while for the family $\Walls{n}$ we only have to compare $S_i$ with the height of the wall
at $i$. For the case of the culminating paths, see below in Section \ref{Sec:Analysis}.
\end{proof}
We now choose the distribution $(p_i)$. Instead of $p_i=1/n$, we will use the probability distribution defined
b
\begin{equation}\label{Eq:Poids}
p_i:=i(2n-i)\kappa_0 +a\quad (\mbox{ for }i=1,\dots,n),
\end{equation}
where
\begin{align*}
\kappa_0&=\frac{3}{2n^2(n+1)}\\
a &=1/4n^3.
\end{align*}
We let the reader check that $(p_i)_{i\leq n}$
is indeed a probability distribution.
The reason for which we use this particular distribution
will appear in the proof of Proposition \ref{Lem:Courbure}.
We will then need the following relation: for
each $1\leq i\leq n-1$,
\begin{equation}\label{Eq:kappa}
p_i-p_{i-1}/2-p_{i+1}/2 = \kappa_0.
\end{equation}
For Algorithm \ref{Algo:CM} to be efficient, we need to know how $S(T)$ is close in law to
$\pi$.
This question is related to the spectral properties of the matrix $P$. In particular, the speed of convergence is governed by the spectral gap ({\em i.e.}\ $1-\lambda$, where $\lambda$ is the largest of the modulus of the eigenvalues different from one,
see \cite{Mix} for example), but this quantity is not known in general.
Some geometrical methods \cite{Dia} allow to bound from below $1-\lambda$, but they assume a precise knowledge of the
structure of the graph defined by the chain $P$. It seems that such results do not apply here.
Instead, we will study the metric properties of the chain $P$ with respect to a natural distance on
$\SsChemins{n}$, and show that
it satisfies a certain \emph{contraction property}.
\subsection{The variant of Algorithm \ref{Algo:CM} for culminating paths}\label{Sec:Analysis}
Unchanged, our Markov chain $P$ cannot generate culminating paths since the path $\mathbf{S}=(a,a,\dots,a)$
would then be isolated: it has no peak/valley
and $\Flip{S}{n}{\downarrow}{-}=(a,a,\dots,-b)$ which is not culminating.
Thus we propose a slight modification for the family $\Culmis{n}$. We only change the definition
of $\Flip{S}{i}{\varepsilon}{\delta}$ when $i=n$ (it won't depend on $\delta$).
Since the maximum is reached at $n$, the $\lceil b/a\rceil +1$ last steps are necessarily
$$
(a,a,\dots,a) \mbox{ or } (-b,a,\dots,a).
$$
We thus define $\Flip{S}{n}{\uparrow}{\delta}$ as the path obtained by changing the $\lceil b/a\rceil +1$ last steps into $(a,a,\dots,a)$ (regardless of their initial values in $\mathbf{S}$)
and $\Flip{S}{n}{\downarrow}{\delta}$ as the path obtained by changing the $\lceil b/a\rceil +1$ last steps into $(-b,a,\dots,a)$.
Notice that despite this change the execution time of each loop of Algorithm \ref{Algo:CM} is still a $\mathcal{O}(1)$:
\begin{itemize}
\item If $I_t< n$, the time needed for the test "$\Flip{S}{I_t}{\varepsilon_t}{\delta_t}$ is in $\SsChemins{n}$" can be considered as constant, since we only have to compare $0,S_i,S_n$.
\item If $I_t=n$, the new value $S_n$ is compared with the maximum of $S$, which can be done in $\mathcal{O}(n)$. Fortunately, this occurs with probability $p_n=\mathcal{O}(1/n)$, so that the time-complexity of each loop is, on average, a $\mathcal{O}(1)$.
\end{itemize}
\section{Error estimates with contraction}\label{Sec:Ricci}
Going back to a more general setting, we consider a Markov chain in a finite set $V$, endowed with a metric $d$.
For a vertice $x\in V$ and a transition matrix $P$, we denote by $P\delta_x$ (resp. $P^t\delta_x$)
the law of the Markov chain associated with $P$ at time $1$ (resp. $t$), when starting from $x$.
For $x,y\in V$, the main assumption made on $P$ is that there is a coupling
between $P\delta_x,P\delta_y$ (that is, a random variable
$(X_1,Y_1)$ with $X_1\stackrel{\mbox{law}}=P\delta_x,Y_1\stackrel{\mbox{law}}=P\delta_y$) such that
\begin{equation}\label{Eq:Courbure}
\mathbb{E}\left[d(X_1,Y_1)\right]\leq (1-\kappa)d(x,y),
\end{equation}
for some $\kappa >0$, which is called the \emph{Ricci curvature} of the chain, by analogy with the Ricci curvature in
differential geometry\footnote{The Ricci curvature is actually the largest positive number such
that \eqref{Eq:Courbure} holds, for all the couplings of $P\delta_x,P\delta_y$ ; here we should rather say
that Ricci curvature is larger than $\kappa$.}.
If the inequality holds, then it implies (\cite{Mix},p.189) that $P$ admits a unique stationary
measure $\pi$
and that, for any $x$,
\begin{equation}\label{Eq:Mixing}
\parallel P^t\delta_x -\pi\parallel_{\mathrm{TV}} \leq (1-\kappa)^t\mathrm{diam}(V),
\end{equation}
where $\mathrm{diam}(V)$ is the diameter of the graph with vertices $V$ induced by the Markov chain.
The notation $\parallel .\parallel_{\mathrm{TV}}$ stands, as usual, for the \emph{Total Variation} distance
over the probability distributions on $V$ defined by
$$
\parallel \mu_1 -\mu_2\parallel_{\mathrm{TV}}:= \sup_{A\subset V} \left|\mu_1(A)-\mu_2(A)\right|.
$$
Hence, a positive Ricci curvature gives the exponential convergence to the stationary measure, with an exact ({\em i.e.}\
\eqref{Eq:Mixing} is non-asymptotic in $t$) bound. In many situations, a smart choice for the coupling between $X_1,X_2$ gives a sharp rate of convergence in eq. \eqref{Eq:Mixing} (see some striking
examples in \cite{Olli}).
\subsection{Metric properties of $P$}
To apply the Ricci curvature machinery, we endow each $\SsChemins{n}$ with the $L^1$-distance
$$
d_1(S,S')=\frac{1}{a+b}\sum_{i=0}^n |S_i-S_i'|.
$$
(Notice that $|S_i-S_i'|$ is always a multiple of $a+b$.) For our purpose, it is fundamental that this metric
space is \emph{geodesic}.
\begin{defi}
A Markov chain $P$ in a finite set $V$ is said to be \emph{geodesic} with respect to the distance $d$
on $V$ if for any two $x,y\in V$ with $d(x,y)=k$, there exist $k+1$ vertices $x_0=x,x_1,\dots,x_k=y$ in $V$ such that for each $i$
\begin{itemize}
\item $d(x_i,x_{i+1})=1$ ;
\item $x_i$ and $x_{i+1}$ are neighbours in the Markov chain $P$ ({\em i.e.}\ $P(x_i,x_{i+1})>0$).
\end{itemize}
This implies in particular that $P$ is irreducible and that the diameter of $P$ is smaller
than $\max_{x,y}d(x,y)$.
\end{defi}
\begin{lem}\label{Lem:Geodesique}
For each family $\Culmis{n}$,$\Walls{n}$,$\Exc{n}$,$\Meandres{n}$ the Markov chain of Algorithm \ref{Algo:CM} is geodesic with respect to $d_1$.
\end{lem}
\begin{proof}[Proof of Lemma \ref{Lem:Geodesique}]
The proof goes by induction on $k$. We fix $S\neq T$ (and denote by $s,t$ the corresponding words) ; we
want to decrease $d_1(S,T)$ by one, by applying the operator $\phi$ with proper $i,\varepsilon$.
We denote by $i_0\in\set{1,\dots,n}$ the first index for which $S\neq T$.
For instance we have $T_{i_0}=S_{i_0}+a+b$. Let $j$ be the position of the left-most peak in $T$ in $\set{i_0+1,i_0+2,\dots,n}$, if such a peak exists.
Then $S':=\Flip{T}{j}{\downarrow}{\delta}$ is also in $\SsChemins{n}$: it is immediate for the
families $\Meandres{n},\Walls{n},\Culmis{n},\Exc{n}$.
We have $d_1(S,S')=k-1$.
If there is no peak in $T$ after $i_0$, then $(t_{i_0+1},t_{i_0+2},\dots,t_n)=(a,a,\dots,a)$. Hence we try to
increase the final steps of $S$ by one.
To do so, we choose $S':=\Flip{S}{n}{\uparrow}{\delta}$ if $S\neq \Flip{S}{n}{\uparrow}{\delta}$,
or $S'=\Flip{S}{j}{\uparrow}{\delta}$ where $j$ is the position of the right-most $-b$ otherwise
(we choose the right-most one to ensure that $\Flip{S}{j}{\uparrow}{\delta}$ remains culminating in the case where
$\SsChemins{n}=\Culmis{n}$.).
\end{proof}
For meanders, excursions and walls, we will show that the Ricci curvature of $P$ with respect to
the distance $d_1$ is (at least) of order $1/n^3$.
\begin{prop}\label{Lem:Courbure}
For the three families $\Meandres{n},\Exc{n},\Walls{n}$, the Ricci curvature of the associated Markov chain,
with weights $(p_i)$ defined as in \eqref{Eq:Poids}, is larger than $\kappa_0$.
\end{prop}
\begin{proof}[Proof of Proposition \ref{Lem:Courbure}]
Fix $\mathbf{S},\mathbf{T}$ in $\SsChemins{n}\in\set{\Meandres{n},\Exc{n},\Walls{n}}$,
we first assume that $\mathbf{S},\mathbf{T}$ are neighbours, for instance $\mathbf{T}=\Flip{S}{i}{\uparrow}$
for some $i$.
\begin{center}
\includegraphics[width=35mm]{Voisins}
\end{center}
Let $(\mathbf{S}^1,\mathbf{S}^2)$ be the random variable in $\SsChemins{n}\times\SsChemins{n}$
whose law is defined by
$$
(\mathbf{S}^1,\mathbf{S}^2)\stackrel{\mbox{(law)}}{=} \left(\phi(\mathbf{S},\mathcal{I},{E}),\phi(\mathbf{T},\mathcal{I},E)\right),
$$
where $\mathcal{I}$ is a r.v. taking values in $\set{1,\dots,n}$ with distribution $\mathbf{p}$ and
$E$ is uniform in $\set{\uparrow,\downarrow}$.
In other words, we run one loop of Algorithm \ref{Algo:CM} simultaneously on both paths.
We want to show that $\mathbf{S}^1,\mathbf{S}^2$ are, on average, closer than $\mathbf{S},\mathbf{T}$.
Different cases may occur, depending on $\mathcal{I}$ and on the index $i$ where $\mathbf{S},\mathbf{T}$ differ.
\noindent{\bf \underline{Case 1. $i=1,2,\dots,n-2$.}\ \ }
\par
\noindent{\bf \underline{Case 1a. $\mathcal{I}=i$.}\ \ }
This occurs with probability $p_i$ and, no matter the value of $E$, we have $\mathbf{S}^1=\mathbf{S}^2$.
\par
\noindent{\bf \underline{Case 1b. $\mathcal{I}=i-1$ or $i+1$.}\ \ }
We consider the case $i-1$. Since $\mathbf{S}$ and $\mathbf{T}$ coincide
everywhere but in $i$, we necessarily have one of these two cases:
\begin{itemize}
\item there is a peak in $\mathbf{S}$ at $i-1$ and neither a peak nor a valley in $\mathbf{T}$ at $i-1$ (as in the figure on the right) ;
\item there is a valley in $\mathbf{T}$ at $i-1$ and neither a peak nor a valley in $\mathbf{S}$ at $i-1$.
\end{itemize}
In the first case for instance, then we may have $d_1(\mathbf{S}^1,\mathbf{S}^2)=2$ if $E=\downarrow$, while
the distance remains unchanged if $E=\uparrow$. The case $\mathcal{I}=i+1$ is identical.
This shows that with a probability smaller than
$p_{i-1}/2+p_{i+1}/2$ we have $d_1(\mathbf{S}^1,\mathbf{S}^2)=2$.
\par
\noindent{\bf \underline{Case 1c. $\mathcal{I}\neq i-1,i,i+1$ and $\mathcal{I}\neq n$.}\ \ }
In this case, $\mathbf{S}$ and $\mathbf{T}$ are possibly modified in $\mathcal{I}$, but if there is
a modification it occurs in both paths. It is immediate since for the families
$\Meandres{n}$,$\Walls{n}$ and $\Exc{n}$ since the constraints are local.
\par
\noindent{\bf \underline{Case 2. $i=n-1$.}\ \ }
In this case, it is easy to check that, because of our definition of $\Flip{S}{n}{\varepsilon}{\delta}$,
we have
$$
\mathbb{E}\left[d_1(\mathbf{S}^1,\mathbf{S}^2)\right]
\leq 1-p_{n-1}+p_{n-2}/2+p_{n}/2 =1-\kappa_0.
$$
\par
\noindent{\bf \underline{Case 3. $i=n$.}\ \ }
We have
$$
\mathbb{E}\left[d_1(\mathbf{S}^1,\mathbf{S}^2)\right]
\leq 1+p_{n-1}/2-p_{n}/2=1-\kappa_0.
$$
\vspace{3mm}
Thus, we have proven that when $\mathbf{S},\mathbf{T}$ only differ at $i$
\begin{align}
\mathbb{E}\left[d_1(\mathbf{S}^1,\mathbf{S}^2)\right]
&\leq 2\times(p_{i-1}/2+p_{i+1}/2) +0\times p_i+1\times(1-p_i-p_{i-1}/2-p_{i+1}/2)\label{Eq:Accroissement}\\
&\leq (1-\kappa_0)\times 1=(1-\kappa_0)d_1(S,T)\notag.
\end{align}
What makes Ricci curvature very useful is that if this inequality
holds for pairs of neighbours then it holds for
any pair, as noticed in \cite{Bub}. Indeed, take $k+1$ paths $S_0=S,S_1,\dots,S_k=T$ as in Lemma
\ref{Lem:Geodesique} and apply the triangular inequality for $d_1$:
\begin{align*}
\mathbb{E}\left[d_1(\phi(S,\mathcal{I},{E}),\phi(T,\mathcal{I},E))\right]
&\leq \sum_{i=0}^{k-1} \mathbb{E}\left[d_1(\phi(S_i,\mathcal{I},{E}),\phi(S_{i+1},\mathcal{I},E))\right]\\
&\leq (1-\kappa_0)k=(1-\kappa_0)d_1(S,T).
\end{align*}
\end{proof}
\begin{rem}
It is easy to exhibit some $S,T$ such that ineq. \eqref{Eq:Accroissement} is in fact
an equality. In
the case where $p_i=1/n$, this equality reads $\mathbb{E}\left[d_1(\mathbf{S}^1,\mathbf{S}^2)\right]=d_1(S,T)$,
and we cannot obtain a positive Ricci curvature (though this does not
prove that there is not another coupling or another distance for which we could
get a $\kappa >0$ in the case $p_i=1/n$.).
\end{rem}
We recall that for each family $\SsChemins{n}$, $\mathrm{diam}(\SsChemins{n})= \max d_1(\mathbf{S},\mathbf{T})
\leq n(n+1)/2$.
Hence, combining Proposition \ref{Lem:Courbure}with Eq. \eqref{Eq:Mixing} gives our main result:
\begin{theo}\label{Th:Mix}
For meanders, excursions and path with walls, Algorithm \ref{Algo:CM} returns an almost uniform sample
of $\pi$, as soon as $T \gg n^3$. Precisely, for any itinialization of Algorithm \ref{Algo:CM},
$$
\parallel \mathbf{S}(T) -\pi\parallel_{\mathrm{TV}} \leq \mathrm{diam}(\SsChemins{n})(1-\kappa)^T
\leq \frac{n(n+1)}{2}\exp\left(-\frac{3}{2n^2(n+1)}T\right).
$$
\end{theo}
Another formulation of this result is that the mixing time of the associated Markov chain, defined
as usual by
\begin{equation}\label{Eq:tmix}
t_{\mbox{mix}}:=\set{\inf\ t\geq 0\ ;\ \sup_{v\in V} \parallel P^t\delta_v -\pi\parallel_{\mathrm{TV}}\leq e^{-1}}
\end{equation}
($e^{-1}$ is here by convention), is smaller than $n^2(n+1)\log n$.
For culminating paths, the argument of Case 1c fails and \eqref{Eq:Accroissement} does not hold, we are not
able to prove such a result as Theorem \ref{Th:Mix}. However, it seems empirically that the mixing time is also of order
$n^3\log n$ (with a constant strongly dependent on $a,b$). A way to prove this could be the following observation: take
$(\mathbf{S}^0,\mathbf{T}^0)=(\mathbf{S},\mathbf{T})$
two any culminating paths, and define
$$
(\mathbf{S}^{t+1},\mathbf{T}^{t+1})=(\phi(\mathbf{S}^t,I_t,\varepsilon_t,\delta_t),\phi(\mathbf{T}^t,I_t,\varepsilon_t,\delta_t)),
$$
where $I_t,\varepsilon_t,\delta_t$ are those in Algorithm \ref{Algo:CM}. The sequence
$
\left(\parallel \mathbf{S}^t-\mathbf{T}^t\parallel_\infty \right)_t
$
is decreasing throughout the process. Unfortunately we cannot get a satisfactory bound for the time needed
for this quantity to decrease by one.
\subsection{Related works}\label{Sec:Related}
Bounding mixing times via a contraction property over the transportation metric is quite a standard technique, the main ideas dating back to Dobrushin (1950's). A modern introduction is made in \cite{Mix}.
For geodesic spaces, this technique has been developped in \cite{Bub} under the name
\emph{path coupling}.
As mentioned in the introduction, the Markov chain $P$ on lattice paths with uniform weights
$p_i=1/n$ has in fact already been
introduced for paths starting and ending at zero (sometimes called \emph{bridges})
in \cite{LRS}, and its mixing time has been estimated in \cite{Wilson}.
Wilson also proves a mixing time of order $n^3\log n$, by showing that \eqref{Eq:Courbure}
holds with a different distance (namely, a kind of Fourier transform of the heights of the
paths)\footnote{Notice that $a,b$ do not have the same meaning in Wilson's paper:
$a$ (resp. $b$) stands for the number of positive (resp. negative) steps.}.
This is the concavity of this Fourier transform which gives a good mixing time, exactly as the concavity
of our $p_i$'s speeds up the convergence of our chain.
Wilson's method is developped only
for bridges in \cite{Wilson} and it is not completely straightforward to use it when the endpoints
are not fixed. For instance, take $n=7$ and $a=b=1$, and consider the paths $+++--++$ and $---++--$. There are more
"bad moves" (moves that take away these paths) than "good moves".
\section{\emph{Coupling From The Past} with $P$}\label{Sec:ProppWilson}
Propp-Wilson's Coupling From The Past (CFTP) \cite{PW} is a very general procedure for the exact sampling of the stationary distribution of a Markov chain. It is efficient if the chain is monotonous with respect to
a certain order relation $\preceq$ on the set $V$ of vertices, with two extremal points
denoted $\hat{0},\hat{1}$ ({\em i.e.}\ such that $\hat{0}\preceq x\preceq\hat{1}$ for any vertex $x$). This is the case
here for each family $\Culmis{n}$,$\Walls{n}$,$\Exc{n}$,$\Meandres{n}$ , with the partial order
$$
\mathbf{S}\preceq \mathbf{T} \mbox{ iff } S_i\leq T_i \mbox{ for any }i.
$$
For the family $\Meandres{10}$ with $a=1,b=-2$ for instance, we have
\begin{align*}
\hat{0}=\hat{0}_{\mbox{meanders}}&=(1,1,-2,1,1,-2,1,1,-2,1),\\
\hat{1}=\hat{1}_{\mbox{meanders}}&=(1,1,1,1,1,1,1,1,1,1).
\end{align*}
We describe CFTP, with our notations, in Algorithm \ref{Algo:CFTP}.
\begin{algorithm}
\caption{CFTP: Exact sampling of a path in $\SsChemins{n}$}
\label{Algo:CFTP}
\begin{algorithmic}
\STATE $\mathbf{S}\leftarrow\hat{0}$, $\mathbf{T}\leftarrow\hat{1}$
\STATE $\dots,I_{-2},I_{-1}\leftarrow$ i.i.d. r.v. with law $\mathbf{p}$
\STATE $\dots,\varepsilon_{-2},\varepsilon_{-1}\leftarrow$ i.i.d. uniform r.v. in $\set{\uparrow,\downarrow}$
\STATE $\dots,\delta_{-2},\delta_{-1}\leftarrow$ i.i.d. uniform r.v. in $\set{+,-}$
\STATE $\tau=1$
\REPEAT
\STATE $\mathbf{S}\leftarrow\hat{0}$, $\mathbf{T}\leftarrow\hat{1}$
\FOR{$t=-\tau$ to $0$}
\STATE {\bf if }$\Flip{S}{I_t}{\varepsilon_t}$ is in $\SsChemins{n}$ {\bf then } $\mathbf{S}\leftarrow \Flip{S}{I_t}{\varepsilon_t}{\delta_t}$
\STATE {\bf if }$\Flip{T}{I_t}{\varepsilon_t}$ is in $\SsChemins{n}$ {\bf then } $\mathbf{T}\leftarrow \Flip{T}{I_t}{\varepsilon_t}{\delta_t}$
\ENDFOR
\STATE $\tau\leftarrow 2\tau$
\UNTIL{$\mathbf{S}=\mathbf{T}$}
\end{algorithmic}
\end{algorithm}
We refer to (\cite{Hagg},Chap.10) for a very clear introduction to CFTP, and we only outline here the reasons why this
indeed gives an exact sampling of the stationary distribution.
\begin{itemize}
\item The output of the algorithm (if it ever ends!) is the state of the chain $P$ that has been running "since time $-\infty$",
and thus has reached stationnarity.
\item The exit condition $\mathbf{S}=\mathbf{T}$ ensures that it is not worth running the chain from $T$ steps earlier,
since the trajectory of any lattice path $\hat{0}\preceq \mathbf{R}\preceq\hat{1}$ is "sandwiched" between those of
$\hat{0},\hat{1}$, and therefore ends at the same value.
\end{itemize}
\begin{figure}[h]
\begin{center}
\includegraphics[width=65mm]{ProppWilson.eps}
\caption{A sketchy representation of CFTP : trajectories starting from $\hat{0},\hat{1}$ at time
$-T/2$ don't meet before time zero, while those starting at time $-T$ do.}
\label{Fig:ProppWilson}
\end{center}
\end{figure}
\begin{prop}\label{Prop:CFTP}
Algorithm \ref{Algo:CFTP} ends with probability $1$ and returns an exact sample of the uniform
distribution over $\SsChemins{n}$. For the families $\Walls{n}$,$\Exc{n}$,$\Meandres{n}$,
this takes on average $\mathcal{O}(n^3(\log n)^2)$ time units.
\end{prop}
Let us mention that in the case where the mixing time is not rigorously known, Algorithm \ref{Algo:CFTP} (when it ends)
outputs an exact uniform sample and therefore is of main practical interest compared to MCMC.
\begin{proof}[Proof of Proposition \ref{Prop:CFTP}]
It is shown in \cite{PW} that Algorithm \ref{Algo:CFTP} returns an exact sampling
in $\mathcal{O}(t_{\mbox{mix}}\log H)$ runs of the chain, where $t_{\mbox{mix}}$ is defined in \eqref{Eq:tmix}
and $H$ is the length of the
longest chain of states between $\hat{0}$ and $\hat{1}$. It is a consequence of the proof of Lemma \ref{Lem:Geodesique}
that $H=\mathcal{O}(n^2)$. We have seen that $t_{\mbox{mix}}=\mathcal{O}(n^3\log n)$. (Recall that each test in
Algorithm \ref{Algo:CFTP} takes, on average, $\mathcal{O}(1)$ time units.)
\end{proof}
We recall that CFTP has a major drawback compared to MCMC. For the algorithm to be correct, we have to reuse
the same random variables $I_t,\varepsilon_t,\delta_t$, so that space-complexity is in fact linear in $n^3(\log n)^2$.
This may become
an issue when $n$ is large.
\section{Concluding remarks and simulations}
\noindent{\bf 1.} In Fig.\ref{Fig:Simus}, we show simulations of the three kinds of paths,
for $a=1,b=2,n=600$.
We observe that the final height of the culminating path is very low (about $30$), it would
be interesting to use our algorithm to investigate the behaviour of this height when $n\to\infty$ ; this question was left open
in \cite{MBM}.
\vspace{4mm}
\begin{figure}[h!]
\begin{center}
\includegraphics[width=140mm]{SimuCulmi600.eps}\\
\includegraphics[width=140mm]{SimuPositif600.eps}\\
\includegraphics[width=140mm]{SimuWall600.eps}\\
\caption{(Almost) uniform paths of length $600$, with $a=1,b=2$. From top to bottom:
a culminating path, a meander, a path with wall (shown by an arch).}
\label{Fig:Simus}
\end{center}
\end{figure}
\noindent{\bf 2.} One may wonder to what extent this work applies to other families $\SsChemins{n}$
of paths. The main assumption is that the family of paths should be a geodesic space w.r.t. distance $d_1$.
This is true for example if the following condition on $\SsChemins{n}$ is fulfilled:
$$
\left(R,T\in\SsChemins{n} \mbox{ and }R\preceq S\preceq T\right) \Rightarrow S\in\SsChemins{n}.
$$
Notice however that this is quite a strong requirement, and it is not verified for culminating paths for instance.
\vspace{4mm}
\noindent{\bf 3.} A motivation to sample random paths is to
make and test guesses for some functionals of these paths, taken on average over $\SsChemins{n}$.
Consider a function $f:\SsChemins{n}\to\mathbb{R}$, we want an approximate value of
$\pi(f):=\mathrm{card}(\SsChemins{n})^{-1}\sum_{s\in\SsChemins{n}}f(s)$, if the exact value is out of reach by calculation. We estimate this quantity by
\begin{equation}\label{Eq:Chapeau}
\hat{\pi}(f):=\frac{1}{T} \sum_{t=1}^T f \left(\mathbf{S}(t)\right),
\end{equation}
(recall that $S(t)$ is the value of the chain at time $t$).
For Algorithm \ref{Algo:CM} to be efficient in practice, we have to bound
\begin{equation}\label{Eq:AMajorer}
\mathbb{P}\left(\left|\pi(f)-\hat{\pi}(f)\right|>r\right),
\end{equation}
for any fixed $r>0$, by a non-asymptotic (in $T$) quantity. This can be done with (\cite{JouOlli}, Th.4-5),
in which one can find concentration inequalities for \eqref{Eq:AMajorer}. The sharpness of these
inequalities depends on $\kappa$ and on the geometrical structure of $\SsChemins{n}$.
\vspace{4mm}
\noindent{\bf Aknowledgements.} Many thanks to Fr\'ed\'erique Bassino and the other members of \textsc{Anr Gamma} for the support ; I also would like to thank \'Elie Ruderman for the English corrections.
A referee raised a serious error in the first version of this paper, I am grateful to them.
|
\section{Introduction}
Gelfand widths are an important concept in classical and modern approximation and complexity theory.
They have found
recent interest in the
rapidly emerging field of compressive sensing \cite{carota06,codade09,do06-2}
because they give general performance bounds
for sparse recovery methods. Since vectors in $\ell_p$-balls, $0<p\leq 1$, can be well-approximated by sparse
vectors, the Gelfand widths of such balls are particularly relevant in this context.
In remarkable papers \cite{ka77,gl84,gagl84} from the 1970s and 80s due to Kashin, Gluskin, and Garnaev, upper and lower estimates for the Gelfand widths of $\ell_1$-balls are provided.
In his seminal paper introducing compressive sensing \cite{do06-2}, Donoho extends these estimates
to the Gelfand widths of $\ell_p$-balls with $p<1$. Unfortunately, his proof of the lower bound contains a gap.
In this article, we address
this issue by supplying a complete proof.
To this end, we proceed in an entirely different way than Donoho.
Indeed, we use compressive sensing methods to establish the lower bound in a more
intuitive way. Our method is new even for the case $p=1$.
For completeness, we also
give a proof of the upper bound based again on compressive sensing arguments. These arguments also provide the same sharp asymptotic behavior for the Gelfand widths of weak-$\ell_p$-balls.
\subsection{Main Result}
In this paper, we consider the finite-dimensional spaces $\ell_p^N$, that is, ${\mathbb{R}}^N$ endowed with the usual $\ell_p$-(quasi-)norm defined, for $x \in {\mathbb{R}}^N$, by
\[
\|x\|_p := \Big( \sum_{\ell=1}^N |x_\ell|^p\Big)^{1/p}, \qquad 0<p<\infty,
\qquad \quad
\|x\|_\infty := \max_{\ell=1,\hdots,N} |x_\ell|.
\]
For $1\leq p \leq \infty$, this is a norm, while for $0<p<1$, it only
satisfies the $p$-triangle inequality
\begin{equation}\label{ptriangle}
\|x+y\|_p^p \leq \|x\|_p^p + \|y\|_p^p,
\qquad x,y \in {\mathbb{R}}^N.
\end{equation}
Thus, $\|\cdot\|_p$ defines a quasi-norm with
(optimal) quasi-norm constant $C = \max\{1,2^{1/p-1}\}$.
As a reminder, $\|\cdot\|_X$ is called a quasi-norm on ${\mathbb{R}}^N$ with quasi-norm constant $C\geq 1$ if it obeys
the quasi-triangle inequality
$$
\|x+y\|_X \leq C(\|x\|_X + \|y\|_X),
\qquad x,y \in {\mathbb{R}}^N.
$$
Other quasi-normed spaces considered in this paper are the spaces weak-$\ell_p^N$,
that is, ${\mathbb{R}}^N$ endowed with the $\ell_{p,\infty}$-quasi-norm defined, for $x \in {\mathbb{R}}^N$, by
$$
\|x\|_{p,\infty} := \max\limits_{\ell = 1,\hdots,N} \ell^{1/p}|x_{\ell}^{\ast}|,
\qquad 0<p \le \infty,
$$
where $x^{\ast}\in {\mathbb{R}}^N$ is a non-increasing rearrangement of $x$.
We shall investigate the Gelfand widths in $\ell_q^N$ of the unit balls
$B_p^N := \{x \in {\mathbb{R}}^N, \|x\|_p \leq 1\}$ and
$B_{p,\infty}^N := \{x \in {\mathbb{R}}^N, \|x\|_{p,\infty} \leq 1\}$ of $\ell_p^N$ and $\ell_{p,\infty}^N$ for $0 < p \le 1$ and $p < q \le 2$.
We recall that the Gelfand width of order $m$
of a subset $K$ of ${\mathbb{R}}^N$
in the (quasi-)normed space $({\mathbb{R}}^N, \|\cdot\|_X)$ is defined as
\[
d^m(K,X) := \inf_{A \in {\mathbb{R}}^{m \times N}} \sup_{v \in K \cap \ker A} \|v\|_X,
\]
where $\ker A := \{v \in {\mathbb{R}}^N, Av = 0\}$ denotes the kernel of $A$.
It is well-known that the above infimum
is actually realized \cite{pi85}.
Let us observe that $d^m(K,X) = 0$ for $m\geq N$
when $0 \in K$, so that we restrict our considerations to the case $m<N$ in the sequel.
Let us also observe that the simple inclusion $B_p^N \subset B_{p,\infty}^N$ implies
$$
d^m(B_p^N,\ell_q^N) \le d^m(B_{p,\infty}^N,\ell_q^N).
$$
From this point on, we aim at finding a lower bound for $d^m(B_p^N,\ell_q^N)$ and an upper bound for $d^m(B_{p,\infty}^N,\ell_q^N)$ with the same asymptotic behaviors.
Our main result is summarized below.
\begin{Theorem}\label{thm:main}
For $0<p\leq 1$ and $p < q \leq 2$, there exist constants
$c_{p,q}, C_{p,q} > 0$ depending only on $p$ and $q$ such that, if $m < N$, then
\begin{equation}\label{upper:lower:Gelfand}
c_{p,q} \min\left\{1,\frac{\ln(N/m)\hspace{-.4mm}+\hspace{-.4mm}1}{m} \right\}^{1/p-1/q}\hspace{-.5mm}
\leq \, d^m(B_{p}^N,\ell_q^N) \,
\leq C_{p,q} \min\left\{1,\frac{\ln(N/m)\hspace{-.4mm}+\hspace{-.4mm}1}{m} \right\}^{1/p-1/q}
\hspace{-.5mm},
\end{equation}
and, if $p<1$,
\begin{equation}
\label{upper:lower:Gelfand2}
c_{p,q} \min\left\{1,\frac{\ln(N/m)\hspace{-.4mm}+\hspace{-.4mm}1}{m} \right\}^{1/p-1/q}\hspace{-.5mm}
\leq d^m(B_{p,\infty}^N,\ell_q^N)
\leq C_{p,q} \min\left\{1,\frac{\ln(N/m)\hspace{-.4mm}+\hspace{-.4mm}1}{m} \right\}^{1/p-1/q}\hspace{-.5mm}.
\end{equation}
\end{Theorem}
In the case $p=1$ and $q=2$, the upper bound of \eqref{upper:lower:Gelfand} with a slightly worse $\log$-term was shown
by Kashin in \cite{ka77} by considering Kolmogorov widths, which are dual to Gelfand widths \cite{lomavo96,pi85}.
The lower bound and the optimal $\log$-term for $p=1$ and $1<q\leq 2$
were provided by Garnaev and Gluskin in \cite{gl84,gagl84}, again via Kolmogorov widths.
An alternative proof of the upper and lower estimates of \eqref{upper:lower:Gelfand} with $p=1$
was given by Carl and Pajor in \cite{capa88}.
They did not pass to Kolmogorov widths, but rather used Carl's theorem
\cite{ca81-1} (see also \cite{cast90,pi85}) that bounds in particular Gelfand numbers from below by entropy numbers,
which are completely understood even for $p,q<1$, see
\cite{sc84-1, guli00, ku01-1}. An upper bound
for $p<1$ and $q=2$ was first provided by Donoho \cite{do06-2} with $\log(N)$ instead of $\log(N/m)$.
With an adaptation of a method from \cite{lomavo96}, Vyb{\'i}ral \cite[Lem. 4.11]{vy08} also provided the upper
bound of \eqref{upper:lower:Gelfand} when $0 < p \leq 1$.
In Section 3, we use compressive sensing techniques to give an alternative proof that provides the upper bound of \eqref{upper:lower:Gelfand2}.
Donoho's attempt to prove the lower bound of \eqref{upper:lower:Gelfand} for the case
$0<p<1$ and $q=2$ consists in
applying Carl's theorem and then using known estimates for entropy numbers,
similarly to the approach by Carl and Pajor for $p=1$.
However, it is unknown whether Carl's theorem extends to
quasi-norm balls, in particular to $\ell_p$-balls with $p<1$. The standard proof of Carl's theorem
for Gelfand widths \cite{cast90, lomavo96} uses duality arguments, which are not available for quasi-Banach spaces. We believe that Carl's theorem
actually fails for Gelfand widths of general quasi-norm balls,
although it turns out to be a posteriori true in our specific situation
due to Theorem \ref{thm:main}.
We shortly comment on the case $q > 2$. Since then $\|v\|_q \leq \|v\|_2$ for all $v \in {\mathbb{R}}^N$, we have the upper estimate
\begin{equation}\label{upperq2}
d^m(B_p^N,\ell_q^N) \leq d^m(B_p^N, \ell_2^N)
\leq C_{p,2} \min\left\{1,\frac{\ln(N/m)+1}{m} \right\}^{1/p-1/2}.
\end{equation}
The lower bound in \eqref{upper:lower:Gelfand} extends to $q>2$, but is unlikely to be optimal in this case. It seems rather
that \eqref{upperq2} is close to the correct behavior. At least for $p=1$ and $q> 2$, \cite{gl82-1} gives lower estimates of
related Kolmogorov widths which then leads to (see also \cite{vy08})
\[
d^m(B_1^N,\ell_q^N) \geq c_q m^{-1/2}\,.
\]
The latter matches \eqref{upperq2} up to the $\log$-factor. We expect a similar behavior for $p<1$,
but this fact remains to be proven.
\subsection{Relation to Compressive Sensing}
Let us now outline the connection to compressive sensing. This emerging theory explores the
recovery of vectors $x \in {\mathbb{R}}^N$ from incomplete linear information $y = Ax \in {\mathbb{R}}^m$, where
$A \in {\mathbb{R}}^{m \times N}$ and $m < N$. Without additional information, reconstruction of $x$ from $y$
is clearly impossible since, even in the full rank case, the system $y = Ax$ has infinitely many solutions.
Compressive sensing makes the additional assumption that $x$ is sparse or at least compressible. A vector $x \in {\mathbb{R}}^N$
is called $s$-sparse if at most $s$ of its coordinates are non-zero. The error of best $s$-term approximation
is defined as
$$
\sigma_s(x)_p := \inf\{ \|x - z \|_p, z \mbox{ is } s\mbox{-sparse} \}.
$$
Informally, a vector $x$ is called compressible if $\sigma_s(x)_p$ decays quickly in $s$.
It is classical to show that, for $q > p$,
\begin{align}
\label{compressible}
\sigma_s(x)_q &\leq \frac{1}{s^{1/p-1/q}} \, \|x\|_p\,,\\
\label{compressible2}
\sigma_s(x)_q & \leq \frac{D_{p,q}}{s^{1/p-1/q}}\, \|x\|_{p,\infty} \,,\qquad \qquad D_{p,q} := (q/p-1)^{-1/q}\,.
\end{align}
This implies that the balls $B_p^N$ and $B_{p,\infty}^N$ with $p\leq 1$ serve as good models for compressible signals:
the smaller $p$, the better $x$ in $B_p^N$ or in $B_{p,\infty}^N$ is approximable in $\ell_q^N$ by $s$-sparse vectors.
The aim of compressive sensing is to find good pairs of linear measurement maps
$A \in {\mathbb{R}}^{m \times N}$ and
(non-linear) reconstruction maps $\Delta : {\mathbb{R}}^m \to {\mathbb{R}}^N$
that recover compressible vectors $x$ with small errors $x-\Delta(Ax)$.
In order to measure the performance of a pair $(A,\Delta)$, one defines,
for a subset $K$ of ${\mathbb{R}}^N$ and a (quasi-)norm $\|\cdot\|_X$ on ${\mathbb{R}}^N$,
\[
E_m(K,X) := \inf_{A \in {\mathbb{R}}^{m \times N}, \, \Delta: {\mathbb{R}}^m \to {\mathbb{R}}^N} \;\; \sup_{x \in K} \|x - \Delta(Ax)\|_X.
\]
Quantities of this type play a crucial role in the modern field of information based complexity \cite{NoWo08}.
In our situation, only linear information is allowed in order to recover $K$ uniformly.
The quantities $E_m(K,X)$ are closely linked to the Gelfand widths, as stated in the following
proposition \cite{do06-2,codade09},
see also \cite{pi86,no95-1}.
\begin{proposition}\label{prop:Em}
Let $K \subset {\mathbb{R}}^N$ be such that $K=-K$ and $K + K \subset C_1 K$ for some $C_1 \geq 2$,
and let $\|\cdot\|_X$ be a quasi-norm on ${\mathbb{R}}^N$ with quasi-norm constant $C_2$.
Note that $C_1 = 2$ if $K$ is a norm ball and that $C_2=1$ if $\|\cdot\|_X$ is a norm.
Then
\[
C_2^{-1} d^m(K,X) \leq E_m(K,X) \leq C_1 d^m(K,X).
\]
\end{proposition}
Combining the previous proposition with Theorem \ref{thm:main}
gives optimal performance bounds for
recovery of compressible vectors in $B_p^N$, $0<p\leq 1$, when the error is measured in
$\ell_q$, $p<q\leq 2$.
Typically, the most interesting case is $q=2$, for which we end up with
\[
c_p \min\left\{1,\frac{\ln(N/m) + 1}{m} \right\}^{1/p-1/2} \leq E_m(B_p^N, \ell_2^N) \leq C_p \min\left\{1,\frac{\ln(N/m)+1}{m}\right\}^{1/p-1/2}.
\]
For practical purposes, it is of course desirable to find matrices $A \in {\mathbb{R}}^{m \times N}$ and efficiently implementable reconstruction maps $\Delta$ that realize the optimal
bound above.
For instance,
Gaussian random matrices $A \in {\mathbb{R}}^{m \times N}$, i.e.,
matrices whose entries are
independent copies of a zero-mean Gaussian variable,
provide optimal measurement maps with high probability \cite{cata06,do06-2,badadewa08}.
An optimal reconstruction map is obtained via basis pursuit \cite{chdosa99,do06-2,cata06}, i.e., via
the $\ell_1$-minimization mapping given by
\[
\Delta_1(y) := {\rm{arg}} \min \|z\|_1 \quad \mbox{ subject to } Az = y.
\]
This mapping can be computed with efficient convex optimization methods \cite{bova04}, and works very well in practice. The proof of the lower bound in \eqref{upper:lower:Gelfand}
will further involve $\ell_p$-minimization for $0<p\le1$
via the mapping
\[
\Delta_p(y) :={\rm{arg}} \min \|z\|_p \quad \mbox{ subject to } Az = y.
\]
A key concept in the analysis of sparse recovery via
$\ell_p$-minimization
is the restricted isometry property (RIP).
This well-established concept in compressive sensing \cite{cata06,carota06-1}
is the main tool for the proof of the upper bound in \eqref{upper:lower:Gelfand2}.
We recall that the $s$th order restricted isometry constant $\delta_s(A)$ of a matrix $A \in {\mathbb{R}}^{m \times N}$ is defined as the smallest $\delta>0$ such that
$$
(1-\delta) \|x\|_2^2 \le \|Ax\|_2^2 \le (1+\delta) \|x\|_2^2
\qquad \mbox{for all $s$-sparse } x \in {\mathbb{R}}^N.
$$
Small restricted isometry constants imply stable recovery by $\ell_1$-minimization, as well as by
$\ell_p$-minimization for $0<p < 1$.
For later reference, we state the following result \cite{carota06-1,ca08,fola09}.
\begin{Theorem}\label{thm:l1} Let $0<p\leq 1$.
If $A \in {\mathbb{R}}^{m \times N}$ has a restricted isometry constant
$\delta_{2s} < \sqrt{2}-1$,
then, for all $x \in {\mathbb{R}}^N$,
\begin{equation}
\label{StableRec}
\| x-\Delta_p(Ax)\|_p^p \leq C\sigma_s(x)_p^p,
\end{equation}
where $C>0$ is a constant that depends only on $\delta_{2s}$.
In particular,
the reconstruction of $s$-sparse vectors is exact.
\end{Theorem}
Given a prescribed $0 < \delta < 1$,
it is known \cite{badadewa08,cata06,mepato09}
that, if
the entries of the matrix $A$ are independent copies of a zero-mean Gaussian variable with variance $1/m$,
then there exist constants $C_1, C_2 > 0$ (depending only on $\delta$) such that $\delta_s(A) \le \delta$ holds with
probability greater than $1-e^{-C_2 m}$
provided that
\begin{equation}\label{mRIP}
m \ge C_1 s \ln(eN/s).
\end{equation}
In particular, there exists a matrix $A \in {\mathbb{R}}^{m \times N}$ such that the pair $(A,\Delta_1)$, and more generally $(A,\Delta_p)$ for $0<p\leq 1$, allows stable recovery in the sense of \eqref{StableRec} as soon as the number of measurements satisfies \eqref{mRIP}.
Vice versa, we will see in Theorem \ref{stable}
that the existence of any pair $(A,\Delta)$ allowing such a stable recovery
forces the number of measurements to satisfy \eqref{mRIP}.
Lemma \ref{pajor2}, which is of independent interest, estimates the minimal number of measurements for the pair $(A,\Delta_p)$ to allow exact (but not necessarily stable) recovery of sparse vectors.
Namely, we must have
\begin{equation}
\label{Improve3s}
m \geq c_1ps\ln(N/(c_2 s))
\end{equation}
for some explicitly given constants $c_1,c_2 > 0$.
In the case $p=1$, this result can be also obtained as a consequence of a corresponding lower bound
on neighborliness of centrosymmetric polytopes, see \cite{do05,LiNo06}.
Decreasing $p$ while keeping $N$ fixed shows that
the bound \eqref{Improve3s} becomes in fact irrelevant for small $p$,
since the bound $m \ge 2s$ holds as soon as there exists a pair $(A,\Delta)$ allowing exact recovery of all $s$-sparse vectors, see \cite[Lem. 3.1]{codade09}.
Combining the two bounds, we see that $s$-sparse recovery by $\ell_p$-minimization forces
$$
m\ge C_1 s \, \big( 1+ p\ln(N/(C_2s)) \big),
$$
for some constants $C_1,C_2 > 0$.
Interestingly, if
such an inequality is fulfilled (with possibly different constants $C_1,C_2$)
and if $A$ is a Gaussian random matrix, then the pair $(A,\Delta_p)$ allows $s$-sparse
recovery with high probability, see \cite{chst08}.
We note, however, that exact $\ell_p$-minimization with $p<1$, as a non-convex optimization program, encounters significant difficulties of implementation.
For more information on compressive sensing, we refer to \cite{ca06-1,carota06,cata06,codade09,do06-2,ra09-1}.
\subsection{Acknowledgements}
The first author is supported by the French National Research Agency (ANR) through the project ECHANGE (ANR-08-EMER-006).
The third and fourth author acknowledge support by the Hausdorff Center for Mathematics, University of Bonn.
The third author acknowledges funding through the WWTF project SPORTS (MA07-004).
\section{Lower Bounds}
In this section, we use compressive sensing methods to establish the lower bound in \eqref{upper:lower:Gelfand}, hence the lower bound in \eqref{upper:lower:Gelfand2} as a by-product.
Precisely, we show the following result,
in which the restriction $q \le 2$ is not imposed.
\begin{proposition}
\label{PropLB}
For $0<p\leq1$ and $p < q\leq \infty$,
there exists a constant $c_{p,q}>0$ such that
\begin{equation}\label{main}
d^m(B_p^N,\ell_q^N) \geq c_{p,q}\min\Big\{1,\frac{\ln(eN/m)}{m}\Big\}^{1/p-1/q}\quad,\quad m<N\,.
\end{equation}
\end{proposition}
\noindent
The proof of Proposition \ref{PropLB} involves several auxiliary steps.
We start with a result \cite{grni03,grni07} on the
unique recovery of sparse vectors
via $\ell_p$-minimization for $0<p\leq 1$.
A proof is included for the reader's convenience.
We point out that,
given a subset $S$ of $[N] := \{1,...,N\}$ and a vector $v \in {\mathbb{R}}^N$,
we denote by $v_{S}$ the
vector that coincides with $v$ on $S$ and that
vanishes on the complementary set $S^c := [N]\setminus S$.
\begin{lemma}\label{NSP} Let $0<p\leq 1$ and $N,m,s \in {\mathbb{N}}$ with $m,s< N$. For a matrix
$A \in {\mathbb{R}}^{m\times N}$, the following statements are equivalent.
\begin{itemize}
\item[(a)] Every $s$-sparse vector $x$ is the unique minimizer of $\|z\|_p$ subject to
$Az = Ax$,
\item[(b)] A satisfies the $p$-null space property of order $s$, i.e.,
for every $v\in \ker A\setminus \{0\}$ and every $S \subset [N]$ with $|S|\leq s$,
$$
\|v_S\|_p^p < \frac{1}{2}\|v\|_p^p\,.
$$
\end{itemize}
\end{lemma}
\begin{Proof}
$(a) \Rightarrow (b)$: Let $v\in \ker A\setminus\{0\}$ and $S \subset [N]$ with $|S| \leq s$.
Since $v = v_{S}+v_{S^{c}}$ satisfies $A v =0$,
we have $Av_{S} = A(-v_{S^c})$.
Then, since $v_{S}$ is $s$-sparse, (a) implies
$$
\|v_S\|_p^p < \|-v_{S^{c}}\|_p^p = \|v_{S^{c}}\|_p^p.
$$
Adding $\|v_{S}\|_p^p$ on both sides and using $\|v_{S^c}\|_p^p + \|v_S\|_p^p = \|v\|_p^p$ gives (b).
$(b) \Rightarrow (a)$: Let $x$ be an $s$-sparse vector and let $S:=\operatorname{supp} x$. Let further $z\neq x$ be such that $Az = Ax$.
Then $v := x-z \in \ker A\setminus \{0\}$ and
\begin{equation}\label{eq11}
\|x\|_p^p \leq \|x_S-z_S\|_p^p + \|z_S\|_p^p = \|v_S\|_p^p + \|z_S\|_p^p\,,
\end{equation}
where the first estimate is a consequence of the $p$-triangle inequality \eqref{ptriangle}.
Clearly, (b) implies $\|v_S\|_p^p < \|v_{S^c}\|_p^p$. Plugging this into (\ref{eq11}) and using that
$v_{S^c} = -z_{S^c}$ gives
$$
\|x\|_p^p < \|v_{S^c}\|_p^p + \|z_S\|_p^p = \|z_{S^c}\|_p^p + \|z_S\|_p^p = \|z\|_p^p.
$$
This proves that $x$ is the unique minimizer of $\|z\|_p$ subject to
$Az = Ax$.
\end{Proof}
\noindent
The next auxiliary step is a well-known combinatorial lemma, see for instance
\cite{avno94,bumirave00,grsl80}, \cite[Lem.\ 3.6]{MePaRu05}.
A proof that provides explicit constants is again included for the reader's convenience.
\begin{lemma} \label{Pajor}
Let $N,s\in {\mathbb{N}}$ with $s < N$\,. There exists a family $\mathcal{U}$ of subsets of $[N]$
such that
\begin{itemize}
\item[(i)] Every set in $\mathcal{U}$ consists of exactly $s$ elements.
\item[(ii)] For all $I, J\in \mathcal{U}$ with $I\neq J$, it holds $|I\cap J| < s/2$.
\item[(iii)] The family $\mathcal{U}$ is ``large'' in the sense that
$$
|\mathcal{U}| \geq \Big(\frac{N}{4s}\Big)^{ s/2 }\,.
$$
\end{itemize}
\end{lemma}
\begin{Proof}
We may assume that $s \le N/4$, for otherwise we can take a family $\mathcal{U}$ consisting of just one element.
Let us denote by $\mathcal{B}(N,s)$ the family
of subsets of $[N]$ having exactly $s$ elements.
This family has size $|\mathcal{B}(N,s)| = \binom{N}{s}$.
We draw an arbitrary element $I_1 \in
\mathcal{B}(N,s)$ and collect in a family $\mathcal{A}_1$ all the sets
$J \in \mathcal{B}(N,s)$ such that $|I\cap J| \ge s/2$\,.
Then $\mathcal{A}_1$ has size at most
\begin{equation}\label{eq10}
\sum\limits_{k=\lceil s/2 \rceil}^{s} \binom{s}{k}\binom{N-s}{s-k}
\le 2^s\max\limits_{\lceil s/2 \rceil\leq k\leq s \; \; }\binom{N-s}{s-k}
= 2^s\binom{N-s}{\lfloor s/2 \rfloor},
\end{equation}
the latter inequality holding because $\lfloor s/2 \rfloor \le (N-s)/2$ when $s \le N/2$.
We throw away $\mathcal{A}_1$ and observe that every element in $J \in \mathcal{B}(N,s)\setminus \mathcal{A}_1$ satisfies $|I_1 \cap J| < s/2$\,.
Next we draw an arbitrary element $I_2 \in \mathcal{B}(N,s)\setminus \mathcal{A}_1$, provided that the latter is not empty.
We repeat the procedure, i.e., we define a family $\mathcal{A}_2$ relative to $I_2$ and draw an arbitrary element $I_3 \in \mathcal{B}(N,s)\setminus(\mathcal{A}_1 \cup \mathcal{A}_2)$, and so forth until no more elements are left.
The size of each set $\mathcal{A}_i$ can always be estimated from above by (\ref{eq10}).
This results in a collection $\mathcal{U} = \{I_1,\hdots,I_L\}$ of subsets of $[N]$ satisfying (i) and (ii).
We finally observe that
\begin{eqnarray*}
L & \geq & \frac{\binom{N}{s}}{2^s\binom{N-s}{\lfloor s/2 \rfloor}}
= \frac{1}{2^s}\, \frac{N(N-1)\cdots(N-s+1)}{(N-s)(N-s-1)\cdots(N-s-\lfloor s/2 \rfloor +1)} \, \frac{1}{s(s-1)\cdots(\lfloor s/2 \rfloor + 1)}\\
& \ge &
\frac{1}{2^s} \, \frac{N(N-1)\cdots (N-\lceil s/2 \rceil +1)}{s(s-1) \cdots (s-\lceil s/2 \rceil +1)}
\ge \frac{1}{2^s} \, \Big(\frac{N}{s}\Big)^{\lceil s/2 \rceil} \ge \Big(\frac{N}{4s}\Big)^{s/2 }.
\end{eqnarray*}
This concludes the proof by establishing (iii).
\end{Proof}
\noindent
We now use Lemma \ref{Pajor} for the final auxiliary result,
which is quite interesting on its own.
It gives an estimate of the minimal number of measurements for exact recovery of sparse vectors via $\ell_p$-minimization,
where $0 < p \le 1$.
\begin{lemma}\label{pajor2}
Let $0<p\leq 1$ and $N,m,s \in {\mathbb{N}}$ with $m < N$ and $s < N/2$.
If $A \in {\mathbb{R}}^{m\times N}$ is a matrix such that every $2s$-sparse vector $x$ is a minimizer of $\|z\|_p$ subject to $Az = Ax$,
then
$$
m\geq c_1ps\ln\Big(\frac{N}{c_2s}\Big)\,,
$$
where $c_1 := 1/\ln 9 \approx 0.455$ and $c_2 := 4$\,.
\end{lemma}
\begin{remark} Lemma 2.6 could be rephrased (with modified constants)
by replacing $2s$-sparse vectors, $s \ge 1$, by $s$-sparse vectors, $s \ge 2$.
In the case $s=1$, it is possible for every $1$-sparse vector $x$ to be a (nonunique) minimizer of $\|z\|_1$ subject to $A z = A x$,
yet $m \ge c_1 p \ln(N/c_2)$ fails for all constants $c_1,c_2>0$.
This can be verified by taking $m=1$ and $A = \begin{bmatrix} 1 & 1 & \cdots & 1 & 1 \end{bmatrix}$.\end{remark}
\begin{Proof}
We consider the quotient space
$$
X:={\mathbb{R}}^N/\ker A = \{[x]:= x+\ker A\,,\,x\in {\mathbb{R}}^N\},
$$
which has algebraic dimension $r := \mbox{rank}\,A \leq m$.
It is a quasi-Banach space equipped with
$$
\|[x]\|_{A,p} := \inf\limits_{v\in \ker A} \|x+v\|_p.
$$
Indeed,
a simple computation reveals that $\|\cdot\|_{A,p}$ satisfies the $p$-triangle inequality, i.e.,
$$
\|[x]+[y]\|_{A,p}^{p} \leq \|[x]\|_{A,p}^p + \|[y]\|_{A,p}^p\,.
$$
By assumption,
the quotient map $[\cdot]$ preserves the norm of every $2s$-sparse vector.
We now choose a family $\mathcal{U}$ of subsets of $[N]$
satisfying (i), (ii), (iii) of Lemma \ref{Pajor}.
For a set $I \in \mathcal{U}$, we define an element $x_I \in \ell_p^N$ with $\|x_I\|_p=1$ by
\begin{equation}
\label{DefxI}
x_I := \frac{1}{s^{1/p}}\sum\limits_{i\in I}e_i\,,
\end{equation}
where $(e_1,\ldots,e_N) $ denotes the canonical basis of ${\mathbb{R}}^N$.
For $I,J \in \mathcal{U}, \; I \not= J$, (ii) yields
$$
\|x_I-x_J\|_p^p > \frac{2s-2s/2}{s} = 1\,.
$$
Since the vector $x_I-x_J$ is a $2s$-sparse vector,
we obtain
$$
\|[x_I]-[x_J]\|_{A,p} = \|[x_I-x_J]\|_{A,p}
= \|x_I-x_J\|_p > 1\,.
$$
The $p$-triangle inequality implies that
$\{[x_I] + (1/2)^{1/p}B_X,~ I\in \mathcal{U}\}$ is a disjoint collection of balls included in the ball $(3/2)^{1/p} B_X$,
where $B_X$ denotes the unit ball of $(X, \|\cdot\|_{A,p})$.
Let $\mbox{vol}(\cdot)$ denote a volume
form on $X$, that is a translation invariant measure satisfying
$\mbox{vol}(B_X)>0$ and $\mbox{vol}(\lambda B_X) = \lambda^r\mbox{vol}(\lambda B_X)$ for all
$\lambda>0$ (such a measure exists since $X$ is isomorphic to ${\mathbb{R}}^r$).
The volumes satisfy the relation
$$
\sum_{I \in \mathcal{U}} \mbox{vol} \big([x_I] + (1/2)^{1/p}B_X \big)
\le
\mbox{vol} \big((3/2)^{1/p} B_X \big).
$$
By translation invariance and homogeneity, we then derive
$$
|\mathcal{U}| \, (1/2)^{r/p} \, \mbox{vol} \big(B_X \big) \leq (3/2)^{r/p} \, \mbox{vol} \big(B_X \big)\,.
$$
As a result of (iii), we finally obtain
$$
\Big( \frac{N}{4s} \Big)^{ s/2 } \le 3^{r/p} \le 3^{m/p}.
$$
Taking the logarithm on both sides gives the desired result.
\end{Proof}
Now we are ready to prove Proposition \ref{PropLB}\,.
The underlying idea is that a small Gelfand width would imply $2s$-sparse recovery for $s$ large enough to violate the conclusion of Lemma \ref{pajor2}. \newline
\begin{Proof}
With $c:=(1/2)^{2/p-1/q}$ and $d:= 2c_1 p / (4+c_1) \approx 0.204 \, p$, we are going to prove that
\begin{equation}
\label{LBwithCsts}
d^m(B_p^N,\ell_q^N) \ge c \, \mu^{1/p-1/q},
\quad \mbox{ where }
\mu:= \min\Big\{1,\frac{d\ln(eN/m)}{m}\Big\}.
\end{equation}
The desired result will follow with $c_{p,q} := c \, d^{1/p-1/q}$.
By way of contradiction, we assume that
$d^m(B_p^N,\ell_q^N) < c \, \mu^{1/p-1/q}$.
This implies the existence of a matrix $A \in {\mathbb{R}}^{m \times N}$ such that, for all $v \in \ker A \setminus \{0\}$,
$$
\|v\|_q < c \, \mu^{1/p-1/q} \|v\|_p.
$$
For a fixed $v \in \ker A \setminus \{0\}$,
in view of the inequalities $\|v\|_p \le N^{1/p-1/q} \|v\|_q$
and $ c \le (1/2)^{1/p-1/q}$,
we derive $1 < (\mu N/2)^{1/p-1/q}$, so that $1 \le 1/\mu < N/2$.
We then define $s:=\lfloor 1/\mu \rfloor \ge 1$, so that
$$
\frac{1}{2 \mu} < s \le \frac{1}{\mu}.
$$
Now, for $v \in \ker A \setminus \{0\}$ and $S \subset [N]$ with $|S| \le 2s$, we have
$$
\|v_S\|_p \le (2s)^{1/p-1/q} \|v_S\|_q
\le (2s)^{1/p-1/q} \|v\|_q < c \, (2s\mu)^{1/p-1/q} \|v\|_p \le
\frac{1}{2^{1/p}} \|v\|_p.
$$
This shows that the $p$-null space property of order $2s$ is satisfied.
Hence, Lemmas \ref{NSP} and \ref{pajor2} imply
\begin{equation}
\label{LBform}
m\geq c_1ps\ln\Big(\frac{N}{c_2s}\Big)\,.
\end{equation}
Besides, since the pair $(A,\Delta_p)$ allows exact recovery of all $2s$-sparse vectors, we have
\begin{equation}
\label{LBform2}
m \ge 2 \, (2s) = c_2 s.
\end{equation}
Using \eqref{LBform2} in \eqref{LBform}, it follows that
$$
m \ge c_1ps\ln\Big(\frac{N}{m}\Big)
= c_1ps\ln\Big(\frac{eN}{m}\Big) - c_1ps
>\frac{c_1 p }{2 \mu} \ln\Big(\frac{eN}{m}\Big) - \frac{c_1}{4} m.
$$
After rearrangement, we deduce
$$
m > \frac{2c_1p}{4+c_1} \, \frac{\ln (eN/m)}{\min \big\{ 1, d \ln(eN/m) / m \big\}}
\ge \frac{2c_1p}{4+c_1} \, \frac{\ln (eN/m)}{ d \ln(eN/m) } \,m = m.
$$
This is the desired contradiction.
\end{Proof}
\begin{remark}
When $m$ is close to $N$,
the lower estimate \eqref{LBwithCsts} is rather poor.
In this case,
a nice and simple argument proposed to us by Vyb{\'i}ral gives the improved estimate
\begin{equation}
\label{LBVyb}
d^m(B_p^N,\ell^N_q) \geq \Big(\frac{1}{m+1}\Big)^{1/p-1/q},\qquad m<N\,.
\end{equation}
Indeed, for an arbitrary matrix $A \in {\mathbb{R}}^{m \times N}$, the kernel of $A$ and the $(m+1)$-dimensional space $\{ x \in {\mathbb{R}}^N: x_i=0 \mbox{ for all } i > m+1 \}$ have a nontrivial intersection.
We then choose a vector $v \neq 0$ in this intersection,
and \eqref{LBVyb} follows from
the inequality $\|v\|_p \leq (m+1)^{1/p-1/q}\|v\|_q$.
\end{remark}
We close this section with the important observation that any measurement/reconstruction scheme that provides $\ell_1$-stability requires a number of measurements scaling at least like the sparsity times a $\log$-term.
This may be viewed as a consequence of Propositions \ref{prop:Em} and \ref{PropLB}.
Indeed, fixing $p<1$, the inequalities \eqref{StableRec} and \eqref{compressible}
imply
$$
d^m(B_p^N,\ell_1^N) \le E_m(B_p^N,\ell_1^N)
\le C \sup_{x \in B_p^N} \sigma_s(x)_1 \le \frac{C }{s^{1/p-1}}.
$$
The lower bound \eqref{main} for the Gelfand width then yields, for some constant $c$,
$$
c \, \min\Big\{1,\frac{\ln(eN/m)}{m}\Big\} \le \frac{1}{s}.
$$
We derive either $s \le 1/c$ or $m \ge c s \ln(eN/m)$.
In short, if $s > 1/c$, then $\ell_1$-stability implies $m \ge c s \ln(eN/m)$
--- which can be shown to imply in turn $m \ge c' s \ln(eN/s)$.
We provide below a direct argument that removes the restriction $s > 1/c$.
It uses Lemma \ref{Pajor} and works also for $\ell_p$-stability with $p < 1$.
It borrows ideas from a paper by Do Ba et al. \cite[Thm. 3.1]{doinprwo10},
which contains the case $p=1$ in a stronger non-uniform version.
\begin{Theorem}\label{stable}
Let $N,m,s \in {\mathbb{N}}$ with $m,s<N$.
Suppose that a measurement matrix $A \in {\mathbb{R}}^{m \times N}$ and a reconstruction map $\Delta: {\mathbb{R}}^N \to {\mathbb{R}}^m$ are stable in the sense that, for all $x \in {\mathbb{R}}^N$,
$$
\|x-\Delta(Ax)\|_p^p \le C \sigma_s(x)_p^p
$$
for some constant $C > 0$ and some $0<p\leq 1$.
Then there exists a constant $C'>0$ depending only on $C$ such that
$$
m \ge C' p \, s \ln(eN/s).
$$
\end{Theorem}
\begin{Proof}
We consider again a family $\mathcal{U}$ of subsets of $[N]$ given by Lemma \ref{Pajor}.
For each $I \in \mathcal{U}$,
we define an $s$-sparse vector $x_I$ with $\|x_I\|_p=1$ as in \eqref{DefxI}.
With $\rho:=(2(C+1))^{-1/p}$,
we claim that
$\{A(x_I+ \rho B_p^N), I \in \mathcal{U}\}$ is a disjoint collection of subsets of $A({\mathbb{R}}^N)$, which has algebraic dimension $r \le m$.
Suppose indeed that there exist $I,J \in \mathcal{U}$ with $I \not= J$
and $z,z' \in \rho B_p^N$ such that $A(x_I+z) = A(x_J+z')$.
A contradiction follows from
\begin{eqnarray*}
1 < \|x_I-x_J\|_p^p & \le &
\| x_I + z - \Delta(A(x_I+z))\|_p^p + \|x_J + z' - \Delta(A(x_J+z')) \|_p^p + \|z\|_p^p + \|z'\|_p^p\\
& \le & C \sigma_s(x_I+z)_p^p + C \sigma_s(x_J+z')_p^p + \|z\|_p^p + \|z'\|_p^p\\
& \le & C \|z\|_p^p + C \|z'\|_p^p + \|z\|_p^p + \|z'\|_p^p \le 2(C+1) \rho^p = 1.
\end{eqnarray*}
We now observe that the collection $\{A(x_I+ \rho B_p^N), I \in \mathcal{U}\}$ is contained in $(1+\rho^p)^{1/p} A(B_p^N)$.
As in the proof of Lemma \ref{pajor2},
we use a standard volumetric argument to derive
$$
|\mathcal{U}| \, \rho^r \, \mbox{vol}\big(A(B_p^N)\big) =
\sum_{I \in \mathcal{U}} \mbox{vol}\big(A(x_I+ \rho B_p^N)\big)
\le \mbox{vol}\big((1+\rho^p)^{1/p} A(B_p^N)\big)
= (1+\rho^p)^{r/p} \mbox{vol}\big(A(B_p^N)\big).
$$
We deduce that
$$
\Big( \frac{N}{4s} \Big)^{ s/2 }
\le (\rho^{-p}+1)^{r/p} \le (\rho^{-p}+1)^{m/p} = (2C+3)^{m/p}.
$$
Taking the logarithm on both sides yields
$$
m \ge c p s \ln(N/(4s)),
\qquad \mbox{ with } c:=1/(2 \ln(2C+3)).
$$
Finally, noticing that $m \ge 2s$ because the pair $(A,\Delta)$ allows exact $s$-sparse recovery, we obtain
$$
m \ge c p s \ln(eN/s) - c p s \ln(4e) \ge c p s \ln(eN/s) - \frac{c \ln(4e)}{2} m.
$$
The desired result follows with $C':=(2c)/(2+c \ln(4e))$.
\end{Proof}
\section{Upper Bounds}
In this section, we establish the upper bound in \eqref{upper:lower:Gelfand2},
hence the upper bound in \eqref{upper:lower:Gelfand} as a by-product.
As already mentioned in the introduction,
the bound for the Gelfand width of $\ell_p$-balls was already provided by Vyb{\'i}ral in \cite{vy08}, but the bound for the Gelfand width of weak-$\ell_p$-balls is indeed new.
\begin{proposition}
For $0<p < 1$ and $p < q\leq 2$,
there exists a constant $C_{p,q}>0$ such that
\begin{equation}\label{f10}
d^m(B_{p,\infty}^N,\ell_q^N) \leq C_{p,q}\min\Big\{1,\frac{\ln(eN/m)}{m}\Big\}^{1/p-1/q}\quad,\quad m<N\,.
\end{equation}
\end{proposition}
The argument relies again on compressive sensing methods.
According to Proposition \ref{prop:Em},
it is enough to establish the upper bound for the quantity $E_m(B_{p,\infty}^N,\ell_q^N)$.
This is done in the following theorem, which we find rather illustrative because it shows that, even when $p < 1$, an optimal reconstruction map $\Delta$
for the realization of the number $E_m(B_{p,\infty}^N,\ell_q^N)$ can be chosen to be the $\ell_1$-minimization mapping, at least when $q \geq 1$.
The argument is originally due to Donoho for $q=2$ \cite[Proof of Theorem 9]{do06-2} and can be extended to all $2\geq q >p$.
\begin{Theorem}\label{thm:optdec} For $0<p < 1$ and $p<q\leq 2$, there exists a matrix $A \in {\mathbb{R}}^{m \times N}$ such that, with $r = \min\{1,q\}$,
\[
\sup_{x \in B_{p,\infty}^N} \|x - \Delta_r(Ax)\|_q \leq C_{p,q} \min \Big\{1,\frac{\ln(N/m)+1}{m} \Big\}^{1/p-1/q},
\]
where $C_{p,q}>0$ is a constant that depends only on $p$ and $q$.
\end{Theorem}
\begin{Proof}
Let $C_1$ be the constant in \eqref{mRIP} relative to the RIP
associated with $\delta=1/3$, say. We choose a constant $D>0$ large enough to have
$$
D/2>e,
\qquad
\frac{D/2}{1+\ln(D/2)} > C_1.
$$
We are going to prove that, for any $x \in B_{p,\infty}^N$,
\begin{equation}
\label{LastPf}
\|x - \Delta_r(Ax)\|_q \le C'_{p,q} \min \Big\{1,\frac{D \ln(eN/m)}{m} \Big\}^{1/p-1/q}
\end{equation}
for some constant $C'_{p,q} > 0$. This will imply the desired result with $C_{p,q}:=C'_{p,q} D^{1/p-1/q}$.\\
{\em Case 1:} $\displaystyle{m > D \ln (eN/m)}$.\\
We define $s \ge 1$ as the largest integer smaller than $ m/ (D \ln (eN/m))$, so that
\begin{equation}
\label{DefS}
\frac{m}{2 D \ln(e N/ m)} \le s < \frac{m}{D \ln(e N/ m)}.
\end{equation}
Putting $t=2s$ and noticing that $t/m < 2/D < 1/e$ and that $u \mapsto u \ln(u)$ is decreasing on $ [0,1/e]$, we obtain
$$
m > \frac{D}{2} t \ln(eN/m) = \frac{D}{2} t \ln(eN/t) + \frac{D}{2} \, m\, (t/m) \ln(t/m)
> \frac{D}{2} t \ln(eN/t) - m \, \ln(D/2),
$$
so that
$$
m > \frac{D/2}{1+\ln(D/2)} t \ln(eN/t) > C_1 t \ln(eN/t).
$$
It is then possible to find a matrix $A \in {\mathbb{R}}^{m \times N}$ with $\delta_{t}(A) \le \delta$.
In particular, we have $\delta_{s}(A) \le \delta$.
Now, given $v := x - \Delta_r(Ax) \in \ker A$, we decompose $[N]$ as the disjoint union of sets $S_1,S_2,S_3,\ldots$ of size $s$ (except maybe the last one) in such a way that
$|v_i| \ge |v_j|$ for all $i \in S_{k-1}$, $j \in S_k$, and $k \ge 2$.
This easily implies $\big( \|v_{S_k}\|_2^2 / s \big)^{1/2} \le \big( \|v_{S_{k-1}}\|_r^r/s \big)^{1/r}$, i.e.,
\begin{equation}
\label{CompSkSk-1}
\|v_{S_k}\|_2 \le \frac{1}{s^{1/r-1/2}} \|v_{S_{k-1}}\|_r,
\qquad k \ge 2.
\end{equation}
Using the $r$-triangle inequality, we have
$$
\|v\|_q^r = \Big\| \sum_{k \ge 1} v_{S_k} \Big\|_q^r
\le \sum_{k \ge 1} \| v_{S_k}\|_q^r
\le \sum_{k \ge 1} \big(s^{1/q-1/2} \| v_{S_k}\|_2 \big)^{r}
\le \sum_{k \ge 1} \Big(\frac{s^{1/q-1/2}}{\sqrt{1-\delta}} \| A v_{S_k}\|_2\Big)^r.
$$
The fact that $v \in \ker A$ implies $Av_{S_1} = -\sum_{k \ge 2} A v_{S_k} $.
It follows that
\begin{eqnarray*}
\|v\|_q^r & \le &
\Big(\frac{s^{1/q-1/2}}{\sqrt{1-\delta}} \Big)^r \Big( \sum_{k \ge 2} \| A v_{S_k}\|_2 \Big)^r
+ \Big(\frac{s^{1/q-1/2}}{\sqrt{1-\delta}} \Big)^r \sum_{k \ge 2} \| A v_{S_k}\|_2^r\\
& \le & 2 \Big(\frac{s^{1/q-1/2}}{\sqrt{1-\delta}} \Big)^r \sum_{k \ge 2} \| A v_{S_k}\|_2^r
\le 2 \Big(\sqrt{\frac{1+\delta}{1-\delta}} s^{1/q-1/2}\Big)^r \sum_{k \ge 2} \| v_{S_k}\|_2^r.
\end{eqnarray*}
We then derive, using the inequality \eqref{CompSkSk-1},
$$
\|v\|_q^r \le
2 \Big(\sqrt{\frac{1+\delta}{1-\delta}} \frac{1}{s^{1/r-1/q}}\Big)^{r} \sum_{k \ge 1} \| v_{S_k}\|_r^r .
$$
In view of the choice $\delta=1/3$ and of \eqref{DefS}, we deduce
\begin{equation}
\label{Eq1LastPf}
\|x-\Delta_r(Ax)\|_q \le 2^{1/r} \sqrt{2} \Big( \frac{2D \ln(eN/m)}{m} \Big)^{1/r-1/q} \|x-\Delta_r(Ax)\|_r\,.
\end{equation}
Moreover, in view of $\delta_{2s} \leq 1/3$ and of Theorem \ref{thm:l1}, there exists a constant $C>0$ such that
\begin{equation}
\label{Eq2LastPf}
\|x - \Delta_r(Ax)\|_r \le C^{1/r} \sigma_s(x)_r.
\end{equation}
Finally, using \eqref{compressible2} and \eqref{DefS}, we have
\begin{equation}
\label{Eq3LastPf}
\sigma_s(x)_r \le \, \frac{D_{p,r}}{s^{1/p-1/r}}
\le \, D_{p,r}\Big( \frac{2D \ln(eN/m)}{m} \Big)^{1/p-1/r}\,.
\end{equation}
Putting \eqref{Eq1LastPf}, \eqref{Eq2LastPf}, and \eqref{Eq3LastPf} together,
we obtain, for any $x \in B_{p,\infty}^N$,
$$
\|x - \Delta_r(Ax)\|_q \le C''_{p,q}\Big( \frac{D \ln(eN/m)}{m} \Big)^{1/p-1/q}
=C''_{p,q}\min \Big\{1,\frac{D \ln(eN/m)}{m} \Big\}^{1/p-1/q},
$$
where $C''_{p,q}:=C^{1/r}D_{p,r}2^{1/r+1/2+1/p-1/q}$.\\
{\em Case 2:} $\displaystyle{m \le D \ln (eN/m )}$.\\
We simply choose the matrix $A \in {\mathbb{R}}^{m \times N}$ as $A=0$.
Then, for any $x \in B_{p,\infty}^N$, we have
$$
\|x - \Delta_r(Ax)\|_q = \|x\|_q \le C'''_{p,q} \|x\|_{p,\infty} \le C'''_{p,q},
$$
for some constant $C'''_{p,q}>0$.
This yields
$$
\|x - \Delta_r(Ax)\|_q \le C'''_{p,q} \min \Big\{1,\frac{D \ln(eN/m)}{m} \Big\}^{1/p-1/q}.
$$
Both cases show that \eqref{LastPf} is valid with $C'_{p,q}:=\max\{C''_{p,q},C'''_{p,q}\}$.
This completes the proof.
\end{Proof}
\begin{remark}
The case $p=1$, for which $r=1$, is not covered by our arguments.
Since $\sup_{x\in B^N_{1,\infty}}\sigma_s(x)_1 \asymp \log(N/s)$ the quantity $\sigma_s(x)_1$ cannot be bounded by a constant times $\|x\|_{1,\infty}$ in order to obtain \eqref{Eq3LastPf}. Instead, the additional log-factor $\log(N/m)$ appears on the right-hand side and therefore in the upper estimate of \eqref{upper:lower:Gelfand2} in the case $p=1$. The correct behavior of the Gelfand widths of weak-$\ell_1$-balls does not seem to be known.
Nonetheless, the inequality $\sigma_s(x)_1 \le \|x\|_1$ is always true. This yields the well-known upper estimate for the Gelfand widths of $\ell_1$-balls and hence completes the proof of Theorem \ref{thm:main}.
\end{remark}
|
\section{Introduction}
The intensity of the Ca II K resonance line observed with spectrographs and Lyot-type filters has long served as a diagnostic of the
solar chromosphere and as an indicator of magnetic activity. However, the literature is lacking in long term analysis of the radiative
properties of solar features at this spectral range. Measurements of large-scale phenomena on Ca II K images provide us with observational
constraints for modeling the solar atmosphere, the differential rotation and meridional circulation, and the behavior of the solar dynamo.
Moreover, they give input for improving the calibration and analysis of historical Ca II K observations, which were made at several
observatories for many years.
We investigated the radiative emission of different types of solar features in the spectral range of the
Ca II K line. We analyzed full-disk 2k x 2k observations from the PSPT Precision Solar Photometric Telescope \citep{coulter1994,ermolli1998,rast2008}. The data were obtained
by using three narrow-band interference filters that sample the Ca II K line with different pass bands. Two filters are centered in the line
core, specifically PK-027 and PK-010. The other filter, PKR-011, is centered in the red wing of the line. We measured the intensity and contrast of various solar features, specifically quiet Sun
(inter-network), network, enhanced network, plage, and bright plage (facula) regions. Moreover, we compared the results obtained with
those derived from the numerical synthesis performed for the three PSPT filters with the RH code (Uitenbroek 2001, 2002) on a set of reference
atmosphere models (Fontenla et al. 2007, 2009).
\section{Observations}
We analyzed 157 sets of full-disk observations available on the MLSO PSPT archive for the period June 7th to July 31st 2007.
All the images were pre-processed to apply the standard instrumental calibration \citep{rast2008}. Then, each image was re-sized
and aligned, by using linear interpolation, to a common reference grid in which the solar disk size is constant.
\begin{figure}
\centering{
\includegraphics[width=4.5cm]{criscuoli_fig_1a.eps}
\includegraphics[width=4.5cm]{criscuoli_fig_1c.eps}
\includegraphics[width=4.5cm]{criscuoli_fig_1b.eps}
}
\caption{\footnotesize
Central disk details extracted from the PSPT images obtained with the three filters sampling the Ca II K line
with different pass bands, specifically PK-027 (top), PK-010 (middle), and PKR-011 (bottom).}
\label{fig1}
\end{figure}
Figure \ref{fig1} shows central disk details extracted from the PSPT images obtained with the three filters sampling the Ca II K line
with different pass bands, specifically PK-027 (top), PK-010 (middle), and PKR-011 (bottom).
Visual inspection reveals the same set of disk features is present in all the images, although their contrast and area look different.
The
transmission profiles measured for the three PSPT filters are larger than the
ones measured for a typical Lyot-type birefringent filter and
do not allow to separate the features of the Ca II K line.
\section{Method}
\begin{figure}
\centering{
\includegraphics[width=4.5cm]{criscuoli_fig_2a_k.eps}
\includegraphics[width=4.5cm]{criscuoli_fig_2a_m.eps}
}
\caption{\footnotesize
Example of {\it PK-027} (top panel) observation analyzed in this study and of the corresponding mask image (bottom panel).
This observation was taken on June 7th 2007. The disk features identified on the images
are show with
different colors; blue, light blue, green, and yellow show quiet Sun, network, enhanced network and plage regions; white
and red show penumbra and umbra regions (visible only blowing up the figure). Black
shows off-disk regions.}
\label{fig3}
\end{figure}
We computed synthetic spectra in the range of CaII K with the RH code developed by \citet{uite2001,uite2002}. Our calculations were performed
assuming nLTE and PRD through the set of atmosphere models presented by \citet{font2007} and \citet{font2009}.
These models describe the temperature and density vertical stratification of various solar features: quiet Sun (model B),
network (model D), enhanced network (model F), plage (model H), and bright plage (model P).
We used the SRPM image decomposition method \citep{font2009} to separate the various solar features.
Figure \ref{fig3} shows an example of the PK-027 observation and corresponding mask (bottom)
obtained with the image decomposition. The various disk features are shown in different colors; blue, light blue, green, and yellow show quiet Sun, network, enhanced network and plage regions.
\begin{figure*}
\centering{\includegraphics[width=12cm]{criscuoli_fig_clv1.eps}
}
\caption{\footnotesize
Comparison between the CLV of intensity values measured on PSPT images (symbols)
for the various disk features and those derived from the spectral synthesis (lines) performed for the same filters on the reference atmosphere models. All the values are normalized to the intensity obtained at the disk center.
The results for the various filters are shown with different colors and line-styles as described in the text.
}
\label{fig4}
\end{figure*}
\section{Results}
For each atmosphere model, we calculated with RH the Response Functions (RFs) at the disk center for the various PSPT filters and atmosphere models. These RFs
indicate that the filters sample quite wide ranges of atmospheric heights, but the radiative signals are dominated by the wings
of the Ca II K line, that form at heights lower than 500 km.
These RFs also confirm that the narrow band filters sample higher layers than broader ones; the filters in the line core also sample higher
layers.
Next, for each filter and disk feature, we analyzed the median and the standard deviation of intensity values measured in 50 equal-area annuli
around the solar disk center. Figure \ref{fig4} shows the comparison between the CLV of intensity values measured on PSPT images (symbols)
for the various disk features and those derived from the spectral synthesis (lines) performed for the same filters on the reference atmosphere models. All the values are normalized to the intensity obtained at the disk center.
The results for the various filters are shown with different symbols and line-styles; green diamond and solid line, violet square and dotted line, and red asterisk and dashed line show the results obtained for PK-027, PK-010, and PKR-011, respectively.
We also analyzed the contrast CLV measured on PSPT images for the various disk features. For each disk position and feature, the contrast was defined as the ratio between the intensity value
derived for the feature and the value obtained for the quiet Sun. Finally, we compared the results of our measurements with those
derived from the spectral synthesis.
\section{Conclusions}
We studied the radiative emission of solar features at the Ca~II~K spectral
range. We analyzed moderate resolution PSPT observations taken with interference filters that sample the Ca~II~K range with different pass bands.
We compared the results of our CLV measurements for various disk features
with the output of the spectral synthesis performed on the most recent set of
semi-empirical atmosphere models presented in the literature by Fontenla et al. (2007, 2009).
We found that the Response Functions derived for the three PSPT filters depend only slightly on the filter band pass and reference
atmosphere; the most sensitive are the ones derived for the filters with the narrower-band and reference atmosphere of bright plage (model P).
We also found that the CLV of contrast values measured for the various features observed at the line core with the PK-010 filter are in good
agreement with the outcomes of the spectral
synthesis. However, the CLV of the contrast obtained for the plage regions (model H) identified on our observations have larger decrease toward the limb than the one derived from the corresponding
atmosphere model used in this study.
\begin{acknowledgements}
SC acknowledgs the financial support of the NSO. This work has been partially supported by the ASI and INAF,
within the ASI-ESS and PRIN-INAF-07 contracts,
respectively. The authors thank ISSI for having hosted meetings of the SSI team.
\end{acknowledgements}
\bibliographystyle{aa}
|
\section{Why the highest resolution?}
Sophisticated astrophysical models often predict spectral-line shapes with asymmetries and wavelength shifts (e.g., 3-dimensional hydrodynamics in stellar atmospheres; isotopic constituents of interstellar lines; velocity gradients in regions of line formation; or simply the hyperfine structure of the atomic transitions themselves). However, the confrontation with observations is often limited by blends, lack of suitable lines, imprecise laboratory wavelengths, insufficient spectral resolution and instrumental imperfections. Observational limits can be pushed by averaging many similar lines, thus averaging small random blends and wavelength errors, although the resolution cannot be increased beyond that at which the original data are sampled. Instrumental designs are often limited by the optical interface of high-resolution spectrometers to [very] large telescopes with their [very] large image scales.
\section{Solar and stellar spectra}
Naturally, high-resolution spectroscopy was first explored for the brightest sources such as the Sun and brighter stars. This has also enabled paradigm changes in the analysis of stellar atmospheres in that, contrary to the days of classical one-dimensional homogeneous and static models, it has been realized that it is not possible -- not even in principle -- to infer detailed stellar properties from analyzing observed line parameters alone, no matter how precisely the spectrum would be measured. Any stellar absorption line is built up by great many contributions from a wide variety of temporally variable inhomogeneities across the stellar surface, whose statistical averaging over time and space produce the line shapes and shifts that can be observed in integrated starlight. While it is possible to compute resulting line profiles (with their asymmetries and wavelength shifts) from hydrodynamic models, the opposite is not feasible because the 3-dimensional structure cannot be uniquely deduced from observed line shapes alone (e.g., Asplund 2005).
While hydrodynamic models may predict detailed properties for a hypothetical spectral line, a confrontation with observations may be unfeasible because stellar lines of the desired species, strength, ionization level, etc., might simply not exist, or be unobservable in practice. In real spectra, lines are frequently blended by stellar rotation, by overlapping telluric lines from the terrestrial atmosphere or else smeared by inadequate instrumental resolution, in practice precluding more detailed studies.
While, in the past, it may have been sufficient to merely resolve the presence of lines, and to determine their strengths, finding line asymmetries implies measurements over smaller fractions of each line-width. Any point on a spectral-line bisector (median) is obtained from intensities at two wavelength positions on either side of the line center. To define not only the bisector slope, but also its curvature requires at least some five points, implying at least ten measurements across the line profile. Given a width of a typical photospheric line of, say, 12~km~s$^{-1}$, an ordinary resolution of R~=~100,000 (3~km~s$^{-1}$) can only indicate the general sense of line asymmetry. Both simulated and observed spectra show how the bisectors degrade when the spectral resolutions decrease towards such values (e.g., Dravins \& Nordlund 1990b; Allende Prieto et al.\ 2002; Ram{\'\i}rez et al.\ 2009). The general appearance of line profiles recorded under different resolutions is shown in Fig.~1.
\begin{figure}
\includegraphics[width=8.3cm]{dravins_fig1.eps}
\caption{Individual line profiles in solar spectra recorded at different resolutions. Three representative \ion{Fe}{i} lines are shown: a deep but somewhat blended line in the violet, a clean line in the green, and a weaker line in the red. Left to right: spectra of solar disk center recorded with a grating spectrometer at R~=~$\lambda$/$\Delta\lambda \approx$~700,000 (Delbouille et al.\ 1989); solar disk center (Neckel 1999) and integrated sunlight (Kurucz et al.\ 1984), both with a Fourier-transform spectrometer at R~$\approx$~350,000; and moonlight with a stellar spectrometer at R~$\approx$~80,000 (Bagnulo et al.\ 2003). Vertical lines mark laboratory wavelengths and thus the `naively' expected line positions. (Dravins 2008) }
\label{Fig1}
\end{figure}
\begin{figure}
\includegraphics[width=8.1cm]{dravins_fig2.eps}
\caption{Examples of individual line bisectors (medians), overplotted on the absorption-line profiles, for three representative \ion{Fe}{ii} lines in R~$\approx$~80,000 spectra (Bagnulo et al.\ 2003) of different F-type stars: 68~Eri (F2~V), $\theta$~Scl (F5~V), and $\nu$~Phe (F8~V). The bisector scale (top) is expanded a factor of 10 relative to the wavelength one. (Dravins 2008) }
\label{Fig2}
\end{figure}
\section{Advancing high-resolution spectroscopy}
High spectral resolution alone may not suffice to obtain high-fidelity spectra. The desired line may be blended with stellar or telluric ones; its laboratory wavelength could be uncertain; the data may be noisy, and the line might be distorted due to stellar rotation or oscillations.
Fig.~2 illustrates the difficulty of finding truly `unblended' lines. Visually, all lines selected appear quite clean but comparing the same line between different stars, it is seen that while bisectors often share some common features, the bisector shapes differ strongly among different lines in the same star (where, under similar conditions of line formation, one would expect them to be similar). Thus, the `noise' -- the cause of bisector deviations from a representative mean -- does not originate from photometric errors but instead is largely `astrophysical' in character, caused by blending lines and similar (Dravins 2008). These spectra represent current high-fidelity spectroscopy, taken from the UVES Paranal Project (Bagnulo et al.\ 2003), where data of particularly low noise were recorded at resolution R~$\approx$~80,000 with the UVES spectrometer (Dekker et al.\ 2000) at the ESO VLT {\it{Kueyen}} unit.
\subsection{Averages over similar lines}
The omnipresence of weak blends makes it impossible to extract reliable asymmetries or shifts from any single line only. However, effects of at least slight blends may be circumvented by forming averages over groups of similar lines which can be expected to have similar physical signatures. Fig.~3 shows such average \ion{Fe}{ii} bisectors for both solar disk center (Delbouille et al.\ 1989) and for integrated sunlight (Kurucz et al.\ 1984). The wavelength scale now is absolute, using laboratory data from the \textsc{FERRUM} project (Johansson 2002; Dravins 2008). For solar disk center, 104 lines were found to be clean enough for such averaging, and 93 for integrated sunlight. From such numbers, reasonably well-defined mean bisectors emerge; however, most atomic species have rather fewer lines.
\begin{figure}[]
\centering
\includegraphics[width=6cm]{dravins_fig3.eps}
\caption{ \ion{Fe}{ii} bisectors in spectra of solar-disk center and of integrated sunlight. Each thin curve is the bisector of one spectral line; the thick dashed is their average. Such averaging over many similar lines (which can physically be expected to have similar intrinsic shapes) is required to reduce effects of weak random blends. The smaller bisector curvature for integrated sunlight reflects both intrinsic line-profile changes across the solar disk, and line broadening due to solar rotation. The vertical scale denotes the intensity in units of the spectral continuum, while the dotted line marks the wavelength expected in a classical 1-dimensional model atmosphere, given by the laboratory wavelength, displaced by the solar gravitational redshift. (Dravins 2008)}
\label{Fig2}
\end{figure}
\subsection{Spatially resolved stellar spectroscopy}
A major milestone in stellar astronomy will be the observation of stars not as point sources, but as extended surface objects. Already, disks of some large stars can be imaged with large telescopes and interferometers, with many more resolvable with future facilities.
3-dimensional models predict spectral changes across stellar disks, differing among various stars and indicating their level of surface `corrugation' (Dravins \& Nordlund 1990a). In stars with `smooth' surfaces, convective blueshifts decrease from disk center towards the limb, since vertical velocities become perpendicular to the line of sight, and the horizontal ones contributing Doppler shifts appear symmetric. Stars with surface `hills' and `valleys' show the opposite, a blueshift increasing towards the limb, where one sees approaching (blueshifted) velocities on the slopes of those `hills' that are facing the observer. Further effects appear in time variability: on a `smooth' star, temporal fluctuations are caused by the random evolution of granules. Near the limb of a `corrugated' star, further variability is added since the swaying stellar surface sometimes hides some granules from direct view.
To realize corresponding observations requires spectrometers with integral-field units and adaptive optics on extremely large telescopes or interferometers. Options for two-dimensional imaging include Fourier transform spectrometers. These provide high spectral resolution for extended objects (e.g., Maillard 2005), although their noise properties have hitherto limited their use in point-source observations
\section{Intergalactic wavelength shifts in quasars}
Wavelength shifts analogous to those in convective stellar photospheres may be expected also in intergalactic absorption lines from various metals, as seen in quasar spectra. These lines typically form in the intergalactic medium of galaxy clusters that happen to fall in the line of sight towards the more distant quasar. Such intergalactic gas undergoes non-random convective motions qualitatively analogous to those in stellar photospheres (even if the dynamic timescales, now on order 100 million years, are \textit{very} much longer). Active galactic nuclei are normally located near such cluster cores and are sources of energetic cosmic rays, heating the surrounding plasma which becomes buoyantly unstable, thus driving convective motions within clusters of galaxies (e.g., Chandran 2004, 2005; Reynolds et al.\ 2005; Sharma et al.\ 2009; Vernaleo \& Reynolds 2006).
Corresponding 3-dimensional hydrodynamic models reproduce several features of the observed structures in the intergalactic gas, and should become possible to use as a model input for computing synthetic spectral-line profiles and wavelength shifts. Differential shifts for lines formed within the same cluster, among those with different oscillator strength, ionization level, etc., could then diagnose intergalactic hydrodynamics, even for the distant Universe. For example, lines formed in the deeper gravitational potential closer to the cluster cores would be expected to be slightly more gravitationally redshifted than lines formed further out. Lines of high excitation potential may be expected to predominantly form in hotter gas (where more atoms are in high-excitation states). If these lines are strong, they will predominantly form in the near-side of the clusters, and therefore be seen as more blueshifted (the convective flow of hot gas is mainly radially outward in the cluster, thus approaching and blueshifted on its near-side, while the observable contributions from the receding and redshifted flow on the far side are weakened due to self-absorption in any stronger line). Such shifts must be modeled and segregated from other possible shifts -- perhaps due to cosmologically changing `constants' -- if such other shifts are to be reliably deduced.
Even if such lineshift modeling is not yet available, one can try to estimate plausible orders of magnitude. If the analogy to stellar line formation holds, one could expect wavelength shifts due to the different statistical weighting of line contributions from different inhomogeneities on the order of perhaps 1\% of the `turbulent' line broadening. Since predicted gas-flow velocities are on order 50--100 km~s$^{-1}$, this implies lineshifts on order 500--1000 m~s$^{-1}$. In order to segregate such intrinsic shifts in complex spectra, to map depth structure from multiple line components or perhaps even to map lateral structure from secular time changes, one clearly will need resolving powers of at least 300,000, and ideally even 1,000,000. Further, since the sources are faint while low-noise spectra are required, extremely large telescopes are probably called for.
\begin{figure*}[t!]\centering
\includegraphics[width=15cm]{dravins_fig4.eps}
\caption{ Existing or planned high-resolution spectrometers for the visual range at existing 8-10 m class telescopes. }
\label{Fig3}
\end{figure*}
\section{Spectrometers at the largest telescopes}
Existing or planned high-resolution stellar spectrometers for the visual at current very large telescopes are summarized in Fig.~4. While some specialized instruments (R~$\approx$~10$^{6}$) have been used in particular to measure interstellar lines (e.g., the Ultra-High-Resolution Facility at the Anglo-Australian Telescope; Diego et al.\ 1995), the currently only seriously high-resolution night-time instrument foreseen at any very large telescope is PEPSI (\textit{Potsdam Echelle Polarimetric and Spectroscopic Instrument}) for the Large Binocular Telescope, planned to reach R~=~320,000 (Strassmeier et al.\ 2003). Among the PEPSI science cases (Strassmeier et al.\ 2004), diagnosing 3-dimensional stellar hydrodynamics was examined, noting that studies of bisector curvature indeed require such resolutions. Steffen \& Strassmeier (2007) further discuss the PEPSI `deep spectrum' project for the highest quality spectra for any star other than the Sun.
At the largest telescopes, challenges in highest-resolution spectroscopy stem from the difficulty to optically match realistically-sized grating spectrometers to the large image scales, i.e., to squeeze starlight into a narrow spectrometer entrance slit while avoiding unrealistically large optical elements (Span{\`o} et al.\ 2006). The use of image slicers would limit the number of measurable spectral orders although some remedy could be offered by adaptive optics (Ge et al.\ 2002; Sacco et al.\ 2004).
Pasquini et al.\ (2005) and D'Odorico et al.\ (2007) discuss future spectrometers. High efficiency requires large optics: the two-grating mosaic for PEPSI makes up a 20$\times$80~cm R4 echelle, while ESPRESSO for the VLT combined focus, and CODEX for the E-ELT aim at twice that: 20$\times$160 cm (Pasquini et al.\ 2008; Span{\`o} et al.\ 2008). Nevertheless, resolutions would not reach much above R~=~100,000.
\subsection{Avoiding telluric absorption?}
Even with superlative spectral resolution and photometric precision, spectra are distorted by telluric absorption and emission lines and bands, compromising the spectral fidelity. Actually, in some wavelength regions, the level of contamination makes it doubtful whether it will ever be possible to record truly high-fidelity solar or stellar spectra from the ground.
In some regions, telluric lines from, say, H$_{2}$O, are sharp and well confined in wavelength, and then (at least in principle) can be identified and removed in the analysis (e.g., Hadrava 2006). However, an accurate removal requires the terrestrial lines to be spectrally resolved, and not even the `high' resolution in solar atlases is fully adequate here. Inspection of telluric lines in clean spectral regions of the Kitt Peak Atlas of integrated sunlight (Kurucz et al.\ 1984) shows that these lines are still unresolved (for an example, see $\lambda$~1250.7 nm), instead displaying the characteristic `ringing' instrumental profile of a Fourier transform spectrometer, even at the resolution R~=~500,000.
There are other and more treacherous telluric features, extending over wider wavelength regions. These include diffuse absorptions due to ozone O$_{3}$ and the oxygen dimer [O$_{2}$]$_{2}$ that appear throughout the visual. Ozone produces the atmospheric transmission cutoff in the near ultraviolet, where old stellar spectrograms can actually be used to infer past amounts of terrestrial ozone (Griffin 2005).
Only in a few cases has it been possible to compare high-resolution spectra recorded from above and beneath the atmosphere. Sirius was observed with the Goddard High-Resolution Spectrograph on the Hubble Space Telescope in the ultraviolet up to 320 nm, while ground-based observations reach down to 305 nm. The overlapping interval is very instructive in demonstrating how the apparent continuum in ground-based recordings can locally be wrong by some 10\% due to diffuse telluric absorptions (Griffin 2005, and private comm.).
Kurucz (2005) computed synthetic telluric spectra, including O$_{3}$ and [O$_{2}$]$_{2}$, showing how they affect the highest-resolution solar spectra. The richness of the telluric spectra means that weak lines with depths of perhaps only some percent or less are often superposed onto the flanks or cores of solar ones, making the reduction for their effects awkward or practically impossible (especially since atmospheric extinction is variable in time; Stubbs et al.\ 2007). In fact, for many spectral regions, this sets a signal-to-noise limit on the order of 100 or worse, irrespective of the photometric precision reached. The affected regions can be inspected in a spectral atlas of telluric absorption (R~$\approx$~300,000), derived from solar data by Hinkle et al.\ (2003).
Observations from orbit would eliminate telluric absorptions (although geocoronal emission would still be present), but any space mission for high-resolution stellar spectroscopy would probably be complex, considering not only that a large telescope is needed but also that the Doppler shift induced by the spacecraft motion implies continuous wavelength changes.
\subsection{Accurate wavelength calibration?}
Grating and Fourier transform instruments establish their wavelength scales through different schemes, and each is likely to have its characteristic signatures in its wavelength noise.
Grating spectrometers commonly use an emission-line lamp (thorium, etc.), and thus their wavelength scale depends upon the accuracy of the corresponding laboratory wavelengths and on how well the lamp at the telescope reproduces the laboratory sources. Traditionally, the wavelength noise among individual lines may have been around 100~$\mathrm{m\, s^{-1}}$ but recent calibrations of thorium-argon hollow-cathode lamps have now decreased the internal scatter to some 10 $\mathrm{m\, s^{-1}}$ (Lovis \& Pepe 2007), further permitting a selection of subsets of lines for best calibration (Murphy et al.\ 2007a).
Fourier transform spectrometers have their specific issues. The interferogram is recorded sequentially for its sinusoidal components, between low Fourier frequencies and high ones (how high determines the spectral resolution). However, the realities of detectors and photon noise normally preclude the entire spectrum being measured at once; instead only one piece at a time must be selected by some pre-filter or similar device. The transmission functions of these pre-filters have to be precisely calibrated to enable a correct continuum level to be set. Its placement to better than 1:100 is no trivial task, as illustrated by the efforts to improve the continuum in the Kitt Peak solar atlases by Neckel (1999) and Kurucz (2005).
\subsection{Future high-fidelity spectroscopy?}
Although several parameters limiting highest-fidelity spectroscopy can be identified, their immediate remedies cannot. Avoiding the terrestrial atmosphere requires ambitious space missions, and improving laboratory wavelengths requires dedicated long-term efforts. Possibly, the least difficult part is to improve instrumental calibrations.
Laboratory devices, in particular the laser frequency comb, permit remarkably precise wavelength determinations, and are also envisioned for future astronomical spectrometers (Araujo-Hauck et al.\ 2007; Murphy et al.\ 2007b; Steinmetz et al.\ 2008). However, it is not necessarily the stability of the calibration device that limits the accuracy in astrophysical spectra because it may in addition require identical light paths for the source and its calibration, and without the calibration light contaminating the source spectrum.
The telescope-spectrometer interface is another issue. Analyses of the UVES spectrometer during asteroid observations reveal a noise of typically 10--50~m~s$^{-1}$, apparently caused by a non-uniform and variable illumination in the image projected by the telescope onto the spectrometer entrance slit (Molaro et al.\ 2008). Given the chromatic nature of atmospheric dispersion, the wavelength dependence of such shifts could also mimic line-depth dependences since stronger atomic lines often occur at shorter wavelengths. Although a more uniform illumination is provided by image slicers or fiber-optics feeds, their use becomes awkward if one requires both extended spectral coverage and the use of (very) large telescopes. For non-solar work, the former requires multi-order echelle spectrometers, and the latter implies (very) large image scales. Light from all image slices, projected onto the focal plane, would overlap adjacent echelle orders (precluding the recording of extended spectral regions), and an optical fiber would need to have a large diameter to embrace most starlight, requiring entrance apertures that are too wide for adequate spectral resolution (or else cause severe light losses). As already mentioned, possible solutions that avoid the construction of huge instruments could include spectrometers with adaptive optics.
Some instrumentation planning committees appear to concentrate on maximum light efficiency for future instruments. While of course desirable, it is a fallacy to believe that highest optical efficiency would be crucial to scientific discovery: what is required is {\textit{adequate}} efficiency to reveal novel features in astronomical sources. Still, a grand challenge remains in designing an efficient truly high-resolution (R~$\approx$~10$^{6}$) and high-fidelity spectrometer for future extremely large telescopes!
\
\acknowledgements
This work is supported by The Swedish Research Council and The Royal Physiographic Society in Lund. It used data from the UVES Paranal Observatory Project of the European Southern Observatory (ESO DDT Program ID 266.D-5655). It also used solar spectral atlases obtained with the Fourier Transform Spectrometer at the McMath/Pierce Solar Telescope situated on Kitt Peak, Arizona, operated by the National Solar Observatory, a Division of the National Optical Astronomy Observatories.
|
\section{INTRODUCTION}
Gravitational lensing generally refers to the bending of light rays of
a background light source by a foreground mass. Gravitational
microlensing, on the other hand, traditionally refers to the special case when multiple images are created
but have separations of less than a few milliarcseconds, and hence are unresolved
with current capabilities. Although the idea of the gravitational deflection of
light by massive bodies well predates General Relativity, and can be traced
as far back as Sir Isaac Newton\footnote{See {\em Schneider et al.}
1992 for a thorough recounting of the history of gravitational
lensing.}, the concept of gravitational microlensing appears to be
attributable to Einstein himself. In 1936 Einstein published a paper
in which he derived the equations of microlensing by a foreground star
closely aligned to a background star ({\em Einstein}, 1936). Indeed it
seems that Einstein had been thinking about this idea as far back as
1912 ({\em Renn et al.}, 1997), and perhaps had even hoped to use the
phenomenon to explain the appearance of Nova Geminorum 1912 ({\em
Sauer}, 2008). However, by 1936 he had dismissed the practical
significance of the microlensing effect, concluding that ``there is no
great chance of observing this phenomenon'' ({\em Einstein}, 1936).
Indeed, there is no ``great chance'' of observing gravitational
microlensing. The optical depth to gravitational microlensing, i.e.,
the probability that any given star is being appreciably lensed at any
given time, is of order $10^{-6}$ toward the Galactic bulge, and is
generally similar or smaller for other lines of sight\footnote{The
phenomenon of gravitational lensing of multiply-imaged quasars by
stars in the foreground galaxy that is creating the multiple images of the quasar
is also referred to as microlensing. In
this instance the optical depth to microlensing can be of order unity. However, in
this chapter we are concerned only with gravitational microlensing of
stars within our Galaxy or in nearby galaxies, where the optical depths
to microlensing are always small.}. Thus at least partly due to this low probability,
the idea of gravitational microlensing lay mostly dormant for five
decades after Einstein's 1936 paper (with some notable exceptions,
e.g., {\em Liebes}, 1964; {\em Refsdal}, 1964). It was not until the
seminal paper by {\em Paczynski} (1986) that the idea of gravitational
microlensing was resurrected, and then finally put into
practice with the initiation of several observational searches for
microlensing events toward the Large and Small Magellenic Clouds
and Galactic bulge in the early 1990s ({\em
Alcock et al.} 1993; {\em Aubourg et al.} 1993; {\em Udalski et al.}, 1993).
The roster of detected microlensing events now numbers in the thousands.
These events have been discovered by many different collaborations, toward several
lines of sight including the Magellenic Clouds ({\em Alcock et al.}, 1997, 2000;
{\em Palanque-Delabrouille et al.}, 1998), and M31 ({\em Paulin-Henriksson et al.} 2002;
{\em de Jong et al.}, 2004; {\em Uglesich et al.}, 2004; {\em Calchi Novati et al.}, 2005),
with the vast majority found toward the Galactic bulge ({\em Udalski et al.}, 2000; {\em Sumi et al.}, 2003;
{\em Thomas et al.}, 2005; {\em Hamadache et al.}, 2006) or nearby fields in the
Galactic plane ({\em Derue et al.}, 2000).
\begin{figure*}
\epsscale{1.8}
\plotone{cartoon.eps}
\caption{\small
Left: The images (dotted ovals) are shown for several
different positions of the source (solid circles), along with the primary lens
(dot) and Einstein ring (long dashed circle). If the
primary lens has a planet near the path of one of the images,
i.e. within the short-dashed lines, then the planet
will perturb the light from the source, creating a
deviation to the single lens light curve.
Right: The magnification as a function of
time is shown for the case of a single lens (solid) and accompanying planet
(dotted) located at the position of the X in the left panel.
If the planet was located at the + instead, then there would be no detectable perturbation, and
the resulting light curve would be essentially identical to the solid curve.
}\label{fig:cartoon}
\end{figure*}
Even before the first microlensing events were detected, it was
suggested by {\em Mao \& Paczynski} (1991) that gravitational
microlensing might also be used to discover planetary companions to
the primary microlens stars. The basic idea is
illustrated in Figure \ref{fig:cartoon}. As the foreground star passes close to the
line of sight to the background source, it splits the source into two
images, which sweep out curved trajectories on the sky as the
foreground lens star passes close to the line of sight to the
source. These images typically have separations of order one
milliarcsecond, and so are unresolved. However, the total area of
these images is also larger than the area of the source, and as a
result the background star also exhibits a time-variable
magnification, which is referred to as a microlensing event. If the
foreground star happens to host a planet with projected separation
near the paths of one of these two images, the planet will further
perturb the image, resulting in a characteristic, short-lived
signature of the planet.
{\em Gould \& Loeb} (1992) considered this novel method of detecting
planets in detail, and laid out the practical requirements for an observational
search for planets with microlensing. In particular, they advocated a ``two tier''
approach. First, microlensing events are discovered and alerted real-time by
a single survey telescope that monitors many square degrees of the Galactic bulge
on a roughly nightly basis. Second, these ongoing events are then densely monitored
with many smaller telescopes to discover the short-lived signatures of planetary
perturbations.
The search for planets with microlensing began in earnest in 1995 with
the formation of several follow-up collaborations dedicated to
searching for planetary deviations in ongoing events ({\em Albrow et
al.}, 1998; {\em Rhie et al.} 2000). These initial searches adopted
the basic approach advocated by {\em Gould \& Loeb} (1993): monitoring
ongoing events alerted by several survey collaborations (e.g., {\em
Udalski et al.}, 1994; {\em Alcock et al.} 1996), using networks of
small telescopes spread throughout the southern hemisphere. Although
microlensing planet searches have matured considerably since their
initiation, this basic approach is still used to this day, with the
important modification that current follow-up collaborations tend to
focus on high-magnification events, which individually have higher
sensitivity to planets ({\em Griest \& Safizadeh}, 1998), as discussed
in detail below.
From 1995-2001, no convincing planet detections were made, primarily
because the relatively small number of events being alerted each year
($\sim 50-100$) by the survey collaborations meant that there were
only a few events ongoing at any given time, and often these were poorly suited for
follow-up. Although interesting upper limits were placed on the
frequency of Jovian planets ({\em Gaudi et al.}
2002, {\em Snodgrass et al.}, 2004), perhaps the most important result
during this period was the development of both the theory and practice of
the microlensing method, which resulted in its transformation from a
theoretical abstraction to a viable, practical method of searching for
planets.
In 2001, the OGLE collaboration ({\em Udalski}, 2003) upgraded to a
new camera with a 16 times larger field of view and so were able to
monitor a larger area of the bulge with a higher cadence. As a
result, in 2002 OGLE began alerting nearly an order of magnitude more
events per year than previous to the upgrade. These improvements in
the alert rate and cadence, combined with improved cooperation and
coordination between the survey and follow-up collaborations, led to
the first discovery of an extrasolar planet with microlensing in 2003
({\em Bond et al.} 2004). The MOA collaboration upgraded to a 1.8m
telescope and 2 deg$^2$ camera in 2004 ({\em Sako et al.}, 2008), and
in 2007 the OGLE and MOA collaborations started alerting $\sim 850$
events per year, thus ushering in the ``golden age'' of microlensing
planet searches.
The first detections of exoplanets with microlensing
({\em Bond et al.}, 2004; {\em Udalski et
al.}, 2005; {\em Beaulieu et al.}, 2006; {\em Gould et al.}, 2006;
{\em Gaudi et al.}, 2008a; {\em Bennett et al.}, 2008; {\em Dong et
al.}, 2009; {\em Janczak et al.}, 2009; {\em Sumi et al.}, 2009) have
provided important lessons about the kinds
of information that can be extracted from observed events. When
microlensing planet surveys were first being developed, it was thought
that the primary virtue of the method would be solely its ability to
provide statistics on large-separation and low-mass planets.
Individual detections would be of little interest because
the only routinely measured property of the planets would be the
planet/star mass ratio, and the nature of the host star in any given
system would remain unknown because the microlensing event itself provides
little information about the properties of the host star, and follow-up
reconnaissance would be difficult or impossible due to the large
distances of the detected systems from Earth. In fact, experience with actual
events has shown that much more information can typically
be gleaned from a combination of a detailed analysis of the light
curve, and follow-up, high-resolution imaging. As a result, in most
cases it is possible to infer the mass of the primary lens and the basic physical properties of
planetary system, including the planet mass and physical separation.
In some special cases it is possible to infer considerably more information.
The theoretical foundation of microlensing is highly technical, the
practical challenges associated with searching for planets with
microlensing are great, and the analysis of observed microlensing
datasets is complicated and time-consuming. Furthermore, although it
is possible to learn far more about individual systems than was
originally thought, extracting this information is difficult, and it
is certainly the case that the systems will never be as well
studied as planets systems found by other methods in the solar
neighborhood.
Give the intrinsic limitations and difficulties of the microlensing method,
and might therefore wonder: ``why bother?'' The
primary motivation for going to all this trouble is that microlensing
is sensitive to planets in a region of parameter space that is
difficult or impossible to probe with other methods, namely low-mass
planets beyond the `snow line', the point in the protoplanetary disk
beyond which ices can exist. The snow line plays a crucial role in
the currently-favored model of planet formation. Thus, even the first
few microlensing planet detections have provided important constraints on
planet formation theories. The second generation of
microlensing surveys will potentially be sensitive to multiple-planet
systems containing analogs of all the solar system planets except
Mercury, as well as to free floating planets. Ultimately, when
combined with the results from other complementary surveys,
microlensing surveys can potentially yield a complete picture of the
demographics of essentially {\it all} planets with masses greater than that of Mars. Thus
it is well worth the effort, even given the drawbacks.
\begin{figure*}
\epsscale{2.1}
\plotone{dia.ps}
\caption{\small
Left: The lens (L) at a distance $D_l$
from the observer (O) deflects light from the source
(S) at distance $D_s$ by the Einstein bending angle $\mbox{\boldmath$\hat\alpha_d$}$.
The angular positions of the images $\mbox{\boldmath$\theta$}$ and unlensed source $\mbox{\boldmath$\beta$}$
are related by the lens equation, $\mbox{\boldmath$\beta$}= \mbox{\boldmath$\theta$}- \mbox{\boldmath$\alpha_d$}
= \mbox{\boldmath$\theta$}-(D_s-D_l)/D_s\mbox{\boldmath$\hat\alpha_d$}$. For a point lens,
$\hat\alpha_d = 4GM/(c^2D_l\theta)$. Right:
Relation of higher-order observables, the angular ($\theta_{\rm E}$) and projected ($\tilde r_{\rm E}$)
Einstein radii, to physical characteristics of the lensing system. Adapted
from {\em Gould}, 2000.
\label{fig:dia}
}\end{figure*}
The primary goal of this chapter is to provide a general
introduction to gravitational microlensing and searches for planets
using this method. Currently, the single biggest obstacle to the
progress of microlensing searches for planets is simply a lack of
human power. There are already more observed binary and planetary events
than can be modeled by the handful of researchers in the world with
sufficient expertise to do so. This situation is likely to get worse
when next-generation microlensing planet surveys come on line.
However, because the theoretical foundation of microlensing is quite
technical and the practical implementation of microlensing planet
searches is fairly complex, it can be very difficult and time
consuming for the uninitiated to gain sufficient familiarity with the
field to make meaningful contributions. This difficulty is
exacerbated by the fact that the explanations of the relevant concepts
are scattered over dozens of journal articles. The aim of this
chapter is to partially remedy this situation by providing a reasonably
comprehensive review, which a beginner can use to gain a least a basic
familiarity with the many concepts, tools, and techniques that are
required to actively participate and contribute to current
microlensing planet searches. Readers are encouraged to peruse reviews
by {\em Paczynski} 1996, {\em Sackett} 1999, {\em Mao} 2008,
and {\em Bennett} 2009a for additional background material.
Section \ref{sec:found} provides an overview of the theory of gravitational microlensing searches for
planets. This is the heart of the chapter, and is fairly long and
detailed. Much of this section may not be of
interest to all readers, and some of the material may be too technical
for a beginner to the subject. Therefore, the first subsection
(\ref{sec:basic}) provides a primer on the basic properties and
features of microlensing. Beginners, or casual readers who are only interested in
a basic introduction to the method itself and its features, can read only
this section (paying particular attention to Figures \ref{fig:cartoon}, \ref{fig:dia},
and \ref{fig:ped}), and then skip to Section
\ref{sec:features}. Section \ref{sec:practice} discusses the
practical implementation of planetary microlensing, while
Section \ref{sec:features} reviews the basic advantages and drawbacks of the
method. Section \ref{sec:results} provides a summary
of the results to date, as well as a brief discussion of their
implications. Section \ref{sec:future} discusses future
prospects for microlensing planet searches, and in particular
the expected yields of a next-generation ground-based planet
search, and a space-based mission.
\section{FOUNDATIONAL CONCEPTS AND EQUATIONS}\label{sec:found}
\subsection{Basic Microlensing}\label{sec:basic}
This section provides a general overview of the basic equations,
scales, and phenomenology of microlensing by a point mass, and a brief
introduction to how microlensing can be used to find planets, and how
such planet searches work in practice. It is meant to be
self-contained, and therefore the casual reader who is not interested
in a detailed discourse on the theory, phenomenology, and practice of planetary
microlensing can simply read this section and then skip to Section
\ref{sec:features} without significant loss of continuity.
A microlensing event occurs when a foreground ``lens'' happen to pass
very close to our line of sight to a more distant background
``source.'' Microlensing is a relatively improbable phenomenon, and
so in order to maximize the event rate, microlensing survey are
typically carried about toward dense stellar fields. In particular,
the majority of microlensing planet surveys are carried out toward the
Galactic center. Therefore, for our purposes, the lens is typically a
main-sequence star or stellar remnant in the foreground Galactic disk
or bulge, whereas the source is a main-sequence star or giant
typically in the bulge.
The left panel of Figure \ref{fig:dia} shows the basic geometry of
microlensing. Light from the source at a distance $D_s$ is deflected
by an angle $\mbox{\boldmath$\hat\alpha_d$}$ by the lens at a distance
$D_l$. For a point lens, $\hat\alpha_d= 4GM/(c^2D_l\theta)$, where $M$
is the mass of the lens, and $\theta$ is the angular separation of the
images of the source and the lens on the
sky\footnote{This form for the bending angle can be derived heuristically by
assuming that a photon passing by an object of mass $M$ at a distance
$b\equiv D_l \theta $ will experience an impulse given by the Newtonian acceleration
$GM/b^2$ over a time $2b/c$, thereby inducing a velocity perpendicular to the
original trajectory of $\delta v = (GM/b^2)(2b/c)= 2GM/(bc)$. The deflection
is then $\delta v/c = 2GM/(bc^2)$. The additional factor of two cannot be derived
classically, and arises from General Relativity (see, e.g., {\em Schneider et al.} 1992).}.
The relation between
$\theta$, and the angular separation $\beta$ between the lens and source in the
absence of lensing, is called the {\it lens equation}, and is
given trivially by $\beta = \theta - \alpha_d$. From basic geometry
and using the small-angle approximation, $\hat \alpha_d (D_s-D_l)=
\alpha_d D_s$. Therefore, for a point lens,
\begin{equation}
\beta = \theta - \frac{4GM}{c^2\theta}\frac{D_s-D_l}{D_sD_l}.
\label{eqn:pmlens}
\end{equation}
\begin{figure*}
\epsscale{2.0}
\plottwo{ped.eps}{ped2.eps}
\caption{\small Basic point-mass microlensing.
(Left) All angles are normalized by the angular Einstein ring radius $\theta_{\rm E}$, shown as a dashed circle
of radius $\theta_{\rm E}$. The source (S) is located at
an angular separation of $u=0.2$ from the lens (L). Two images are created, one image outside
the Einstein ring (I$_+$), on the same side of the lens as the source with position from the lens of $y_+=0.5(\sqrt{u^2+4}+u)$, and one
inside the Einstein ring, on the opposite side of the lens as the source with position from the lens
of $y_-=-0.5(\sqrt{u^2+4}-u)$. The images are compressed radially but elongated tangentially. Since surface
brightness is conserved, the magnification of each image is just the ratio of its area to the area of the source.
Since the images are typically unresolved, only the total magnification of the two images is measured,
which depends only on $u$. (Right) Magnification as a function of time (light curves), for the
ten trajectories shown in the left panel with impact parameters $u_0=0.01,0.1,0.2,...,1.0$. Time is relative
to the time $t_0$ of the peak of the event (when $u=u_0$), and in units of the angular Einstein crossing time $t_{\rm E}$.
Higher magnification implies more elongated images, which leads to increased sensitivity to planetary companions.
Adapted from {\em Paczynski} 1996.
}\label{fig:ped}
\end{figure*}
The left panel of Figure \ref{fig:ped} shows the basic source and image
configurations for microlensing by a single
point mass. From Equation \ref{eqn:pmlens}, if the lens is exactly aligned with the
source ($\beta=0$), it images the source into an ``Einstein ring'' with a radius
$\theta_{\rm E}=\sqrt{\kappa M \pi_{\rm rel}}$,
where $M$ is the mass of the lens, $\pi_{\rm rel}={\rm AU}/D_{\rm rel}$
is the relative lens-source
parallax, $D_{\rm rel} \equiv (D_l^{-1}-D_s^{-1})^{-1}$ is the relative lens-source
distance, and $\kappa=4G/c^2{\rm AU} \simeq 8.14~{\rm mas}~M_\odot^{-1}$. It is also
instructive to note that $\theta_{\rm E} = \sqrt{2R_{\rm Sch}/D_{\rm rel}}$, where
$R_{Sch} \equiv 2GM/c^2$ is the Schwarzschild radius of the lens. Quantitatively,
\begin{equation}
\theta_{\rm E} = 550~{\rm mas}~\left(\frac{M}{0.3 M_\odot}\right)^{1/2}
\left(\frac{\pi_{\rm rel}}{125~\mu{\rm as}}\right)^{1/2},
\label{eqn:thetaequant}
\end{equation}
which corresponds to a physical Einstein ring radius at the distance of the lens of,
\begin{eqnarray}
r_{\rm E} &\equiv& \theta_{\rm E}D_l\\
&=&2.2{\rm AU}\left(\frac{M}{0.3 M_\odot}\right)^{1/2}
\left(\frac{D_s}{8~{\rm kpc}}\right)^{1/2}
\left[\frac{x(1-x)}{0.25}\right]^{1/2},
\label{eqn:requant}
\end{eqnarray}
where $x\equiv D_l/D_s$.
Normalizing
all angles on the sky by $\theta_{\rm E}$, we can define $u=\beta/\theta_{\rm E}$ and $y=\theta/\theta_{\rm E}$.
Using these definitions and the definition for $\theta_{\rm E}$, the lens equation (\ref{eqn:pmlens})
reduces to the form
\begin{equation}
u = y - y^{-1}.
\end{equation}
This is equivalent to a quadratic equation in $y$: $y^2-uy-1=0$. Thus in the case of imperfect
alignment ($u \ne 0$), there are two images, with positions
\begin{equation}
y_{\pm}=\pm \frac{1}{2}(\sqrt{u^2+4}\pm u).
\label{eqn:ypmfirst}
\end{equation}
The positive, or ``major'' image is always outside the Einstein ring, whereas the negative,
or ``minor'' image is always inside the Einstein ring\footnote{The images
formed by a gravitational lens can also be found by determining the relative time delay as a function
of the (vector) angle $\mbox{\boldmath$\theta$}$ for hypothetical light rays propagating from the source. This time
delay function includes the effect of the difference in the geometric path length, as well the gravitational
(Shapiro) delay, as a function of $\mbox{\boldmath$\theta$}$. The images are then located at the stationary
points (maxima, minima, or saddle points) of the time delay surface. For a point-mass lens,
the positive solution in Equation \ref{eqn:ypmfirst} corresponds to a minimum in the time delay surface, and
so is also referred to as the ``minimum'' image. The negative solution corresponds
to a saddle point, and so is also referred to as the ``saddle'' image.
There is formally a third image, corresponding to the maximum of the time delay surface, but this
image is located behind the lens, and is infinitely demagnified (for a true point lens).}.
As can be seen in Figure \ref{fig:ped}, the angular separation between these images at
the time of closest alignment is is $\sim 2\theta_{\rm E}$. Thus for typical lens masses ($0.1-1~M_\odot$) and
lens and source distances ($1-10~{\rm kpc}$),
$\theta_{\rm E}\la {\rm mas}$ and so the images are not resolved.
However, the images are also distorted, and since surface brightness is conserved, this implies
they are also magnified. The magnification of each image is just the ratio of the area
of the image to the area of the source. As can be seen in Figure \ref{fig:ped}, the images
are typically elongated tangentially by an amount $y_\pm/u$, but are compressed
radially by an amount ${\rm d}{y_\pm}/{\rm d}u$. The magnifications
are then
\begin{equation}
A_\pm = \left| \frac{y_\pm}{u} \frac{{\rm d}y_\pm}{{\rm d}u} \right| = \frac{1}{2}\left[\frac{u^2+2}{u\sqrt{u^2+4}} \pm 1\right],
\label{eqn:apmped}
\end{equation}
and thus the total magnification is
\begin{equation}
A(u)= \frac{u^2+2}{u\sqrt{u^2+4}}
\label{eqn:atot1}.
\end{equation}
Note that, as $u \rightarrow \infty$, $A_+\rightarrow 1$ and $A_- \rightarrow 0$.
Also, for $u \ll 1$, the magnification takes the form $A \simeq 1/u$.
Because the lens, source, and observer are all in relative motion,
the angular separation between the lens and source is a function of time.
Therefore, the magnification of the source is also a function of time:
a microlensing event. In the simplest case of uniform, rectilinear motion, the
relative lens-source motion can be parametrized by
\begin{equation}
u(t) = \left[u_0^2 + \left(\frac{t-t_0}{t_{\rm E}}\right)^2\right]^{1/2},
\label{eqn:uoft1}
\end{equation}
where $u_0$ is the minimum separation
between the lens and source (in units of $\theta_{\rm E}$), $t_0$ is the
time of maximum magnification (corresponding to when $u=u_0$), and
$t_{\rm E}$ is the time scale to cross the angular Einstein ring radius,
$t_{\rm E} \equiv \theta_{\rm E}/\mu_{\rm rel}$, where $\mu_{\rm rel}$ is the proper motion
of the source relative to the lens. This form for $u$
gives rise to a smooth, symmetric
microlensing event with a characteristic form (often
called a ``Paczynski curve''), as shown
in the right panel of Figure \ref{fig:ped}.
The typical timescales for events toward the Galactic bulge
are of order a month,
\begin{equation}
t_{\rm E} = 19~{\rm d}\left(\frac{M}{0.3 M_\odot}\right)^{1/2}
\left(\frac{\pi_{\rm rel}}{125~\mu{\rm as}}\right)^{1/2}
\left(\frac{\mu_{\rm rel}}{10.5~{\rm mas/yr}}\right)^{-1},
\label{eqn:tequant}
\end{equation}
but can range from less than a day to
years. The magnification is $A>1.34$ for $u\le 1$, and so the
magnifications are substantial and easily detectable.
If the lens star happens to have a planetary companion, and the planet
happens to be located near the paths of one or both of the two images
created by the primary lens, the companion can create a
short-timescale perturbation on the primary microlensing event when
the image(s) sweep by the planet ({\em Mao \& Paczynski} 1991; {\em
Gould \& Loeb}, 1992). There are two conceptually different channels
by which planets can be detected with microlensing, delineated by
the maximum magnification $A_{\rm max}$ of the primary event in which the planet is detected.
Consider the case of a relatively low-magnification
(i.e., $A_{\rm max} \la 10$ or $u_0 \ga 0.1$) primary event, such as illustrated in Figure \ref{fig:cartoon}.
Over the course of
the primary event, the two images sweep out a path on the sky, with
the path of the major (minimum) image entirely contained outside the
Einstein ring radius, and the path of the minor (saddle) image
entirely contained within the Einstein ring radius. In the
main detection channel, the primary hosts a planet which happens to be located near the
path of one\footnote{As can be seen in Figure \ref{fig:cartoon}, for low-magnification events,
the two images are well-separated, and thus a low mass ratio planet will generally only
significantly perturb one of the two images.} of these two images for such a low-magnification event.
As the image sweeps by the position of the planet, the planet will further perturb
the light from this image and yield a short-timescale deviation
({\em Gould \& Loeb}, 1992). The
duration of this deviation is $\sim t_{{\rm E},p} = q^{1/2} t_{\rm E}$,
where $q=m_p/M$ is the mass ratio and $m_p$ is the planet mass, and the magnitude of the perturbation depends on how close the perturbed image
passes by the planet.
Given a range of primary event durations of $\sim 10-100~{\rm days}$,
the duration of the perturbations range from a few hours for an
Earth-mass planet to a few days for Jovian-mass planets. As can
be seen in Figure \ref{fig:dia}, the location of
the perturbation relative to the peak of the primary event depends on
two parameters: $\alpha$, the angle of the projected star-planet axis with respect to the
source trajectory\footnote{Conventionally, both the lensing
deflection angle and the source trajectory angle are denoted by
the symbol $\alpha$. In order to maintain contact with the literature,
this unfortunate convention is maintained here, but to avoid confusion
the deflection angle is denoted with the subscript ``$d$''.},
and $d$, the instantaneous angular separation between the planet and host star in
units of $\theta_{\rm E}$.
Because the orientation of the source trajectory relative to the planet
position is random, the time of this perturbation is not predictable
and the detection probability is $\sim A(t_{0,p}) \theta_{{\rm
E},p}/\theta_{\rm E}$, where $A(t_{0,p})$ is the unperturbed
magnification of the image that is being perturbed at the time
$t_{0,p}$ of the perturbation ({\em Horne et al.}, 2009), and
$\theta_{{\rm E},p} \equiv q^{1/2} \theta_{\rm E}$ is the Einstein ring
radius of the planet. Here the factor $A$ accounts for the fact that
the area of the image plane covered by magnified images is larger by
their magnification.
Since the planet must be located near one of the two primary images in
order to yield a detectable deviation, and these images are always
located near the Einstein ring radius when the source is significantly
magnified (see Figure \ref{fig:cartoon}), the sensitivity of the
microlensing method peaks for planet-star projected separations of
$\sim r_{\rm E}$, i.e., for $d\sim 1$. However, microlensing can also
detect planets well outside the Einstein ring ($d\gg 1$), albeit with less
sensitivity. Since the magnification of the minor image decreases
with position as $y_-^4$ (see Equations \ref{eqn:ypmfirst} and
\ref{eqn:apmped}), microlensing is generally not sensitive to planets
$d\ll 1$, i.e., close-in planets.
Given the existence of a planet with a projected separation within a
factor of $\sim 2$ of the Einstein ring radius, the detection
probabilities range from tens of percent for Jovian planets to a few
percent for Earth-mass planets ({\em Gould \& Loeb}, 1992; {\em
Bolatto \& Falco}, 1994; {\em Bennett \& Rhie}, 1996; {\em Peale},
2001). Detecting planets via the main channel requires substantial
commitment of resources because the unpredictable nature of the
perturbation requires dense, continuous sampling, and furthermore the
detection probability per event is relatively low so many events must
be monitored.
A useful feature of low-magnification planetary microlensing events
such as that shown in Figure \ref{fig:cartoon} is that it is possible
to essentially `read off' the light curve parameters from the observed
features ({\em Gould \& Loeb}, 1992; {\em Gaudi \& Gould}, 1997b).
For the primary event, the three gross observable parameters are the
time of maximum magnification $t_0$, the peak magnification $A_{\rm
max}$, and a measure of the duration of the event such as its
full-width half-maximum $t_{FWHM}$. The latter two observables can
then be related to the Einstein timescale $t_{\rm E}$ and impact parameter
$u_0$ using Equations \ref{eqn:atot1} and \ref{eqn:uoft1}
\footnote{For the purposes of exposition, this discussion ignores
blending, which is generally important and complicates the
interpretation of observed light curves. See Section
\ref{sec:lightcurves}.}. For example, for small $u_0$ we have $u_0
\sim A_{\rm max}^{-1}$ and $t_{\rm E} \sim
0.5t_{FWHM}u_0^{-1}$. Unfortunately, only $t_{\rm E}$ contains any
information about the physical properties of the lens, and then only
in a degenerate combination of the lens mass, distance, and relative
lens-source proper motion. However, as discussed in detail in Section
\ref{sec:properties}, in many cases it is often possible to obtain
additional information which partially or totally breaks this
degeneracy. In particular, in those cases where it is possible to
isolate the light from the lens itself, this measurement can be used
to constrain the lens mass ({\em Bennett et al.}, 2007). The three
parameters that characterize the planetary perturbation are the
duration of the perturbation, the time of the perturbation $t_{0,p}$,
and the magnitude of the perturbation, $\delta_p$. As mentioned
previously, the duration of
the perturbation is proportional to $t_{{\rm E},p}\equiv q^{1/2}t_{\rm E}$, and
so gives the planet/star mass ratio $q$. The time and magnitude of
the perturbation then specify the projected separation $d$ and angle
of the source trajectory with respect to the binary axis $\alpha$,
since $t_{0,p}$ and $\delta_p$ depend on the location of the planet
relative to the path of the perturbed image ({\em Gaudi \& Gould},
1997b).
The other channel by which microlensing can detect planets is in high
magnification events ({\em Griest \& Safizadeh}, 1998)\footnote{There
is no formal definition for `high'-magnification events, however
typically high-magnification refers to events with maximum
magnification $A_{\rm max} \ga 100$, corresponding to impact
parameters $u_0\la 0.01$.}. In addition to perturbing images that
happen to pass nearby, planets will also distort the perfect circular
symmetry of the Einstein ring. Near the peak of high-magnification
events, as the lens passes very close to the observer-source line of
sight (i.e.\ when $u\ll 1$), the two primary images are highly
elongated and sweep along the Einstein ring, thus probing this
distortion. For very high-magnification events, these images probe
nearly the entire Einstein ring radius and so are sensitive to all
planets with separations near $r_{\rm E}$, regardless of their orientation
with respect to the source trajectory. Thus high-magnification events
can have nearly 100\% sensitivity to planets near the Einstein ring
radius, and are very sensitive to low-mass planets ({\em Griest \&
Safizadeh}, 1998). However, these events are rare: a fraction $\sim
1/A_{\rm max}$ of events have maximum magnification $\ga A_{\rm max}$.
Fortunately, these events can often be predicted several hours to
several days ahead of peak, and furthermore the times of high
sensitivity to planets are within a full-width half-maximum of the
event peak, or roughly a day for typical high-magnification events
({\em Rattenbury et al.}, 2002). Thus scarce observing resources can
be concentrated on these few events and only during the times of
maximum sensitivity. Because the source stars are highly magnified,
it is also possible to use more common, smaller-aperture telescopes.
\subsection{Theory of Microlensing}\label{sec:theory}
Gravitational lensing can be thought of as the mapping
$\mbox{\boldmath$\beta$} \rightarrow \mbox{\boldmath$\theta$}$ between the angular
position of a source $\mbox{\boldmath$\beta$}$ in the absence of lensing to the angular position(s)
$\mbox{\boldmath$\theta$}$ of the image(s) of the source under the action of the gravitational lens.
This mapping is given by the lens equation,
\begin{equation}
\mbox{\boldmath$\beta$} = \mbox{\boldmath$\theta$} - \mbox{\boldmath$\alpha_d(\theta)$},
\label{eqn:lenseq}
\end{equation}
where $\mbox{\boldmath$\alpha_d$}$ is deflection of the source due to the lens.
Consider a source at a distance $D_s$ and a lens
located at a distance $D_l$ from the observer.
Figure \ref{fig:dia} shows the lensing geometry. The deflection angle $\mbox{\boldmath$\alpha_d$}$ is related
to the angle $\mbox{\boldmath$\hat \alpha_d$}$ by which the lens mass bends the light ray from the source by
$\mbox{\boldmath$\hat \alpha_d$}(D_s-D_l)=\mbox{\boldmath$\alpha_d$}D_s$.
Assume the lens is a system of $N_L$ point masses
each with mass $m_i$ and position $\mbox{\boldmath$\theta$}_{m,i}$. Further assume
that the lenses are static (or, more precisely, moving much more slowly than $c$), and
their distribution along the line of sight is small in comparison to $D_l$, $D_s$, and $D_s-D_l$.
The deflection angle is then,
\begin{equation}
\mbox{\boldmath$\alpha_d$}(\mbox{\boldmath$\theta$}) = \frac{4G}{D_{\rm rel} c^2} \sum_i^{N_l} m_i
\frac{\mbox{\boldmath$\theta$}- \mbox{\boldmath$\theta$}_{m,i} }{|\mbox{\boldmath$\theta$}-\mbox{\boldmath$\theta$}_{m,i}|^2},
\label{eqn:bend}
\end{equation}
where $D_{\rm rel} \equiv (D_l^{-1}-D_s^{-1})^{-1}$. See {\em Schneider et al.} (1992) and {\em Petters et al.} (2001)
for the expression for $\mbox{\boldmath$\alpha_d$}$ for a general mass distribution,
as well for the derivation of Equations \ref{eqn:lenseq} and \ref{eqn:bend}
from the time delay function and ultimately the metric.
It is common practice to normalize all angles to the angular Einstein ring radius,
\begin{equation}
\theta_{\rm E} \equiv \left(\kappa M \pi_{\rm rel}\right)^{1/2},
\label{eqn:thetae}
\end{equation}
where $M \equiv \sum_i^{N_l}m_i$ is the total mass of the lens,
$\pi_{\rm rel} \equiv {\rm AU}/D_{\rm rel}$ is the relative lens-source parallax,
and $\kappa=4G/(c^2{\rm AU})\simeq 8.14~{\rm mas}~M_\odot^{-1}$.
The reason for this convention is clear when considering the single lens (Section \ref{sec:point}).
The dimensionless vector source
position is defined to be ${\bf u} \equiv \mbox{\boldmath$\beta$}/\theta_{\rm E}$, and the dimensionless vector
image positions are defined to be ${\bf y} \equiv \mbox{\boldmath$\theta$}/\theta_{\rm E}$.
It is often convenient to write the lens equation in complex coordinates ({\em Witt}, 1990).
Defining the components of the (dimensionless) source position to be ${\bf u}=(u_1,u_2)$
and the image position(s) to be ${\bf y}=(y_1,y_2)$, the two-dimensional
source and images positions can be expressed in complex form as,
$\zeta=u_1 + i u_2$ and $z= y_1 + i y_2$. The lens equation can now be rewritten
\begin{equation}
\zeta = z - \sum_i^{N_l} \frac{\epsilon_i}{{\bar z} - {\bar z}_{m,i}},
\label{eqn:clenseq}
\end{equation}
where $z_{m,i}$ is the position of mass $i$, $\epsilon \equiv m_i/M$.
The overbars denote complex conjugates, which in Equation \ref{eqn:clenseq}
arise from the identity $z/|z|^2 \equiv {\bar z}^{-1}$.
This equation can then be solved to find the image positions $z_j$.
Lensing conserves surface brightness, but because of the mapping the
angular area of each image of the source is not necessarily equal to
angular area of the unlensed source. Thus the flux of each image (the
area times the surface brightness) is different from the flux of the
unlensed source: the source is magnified or demagnified. For a small
source, the magnification $A_j$ of each image $j$ is given by the
amount the source is ``stretched'' due to the lens mapping.
Mathematically, the amount of stretching is given by the inverse of the determinant of the Jacobian of the mapping
(\ref{eqn:clenseq}) evaluated at the image position,
\begin{equation}
A_j = \left. \frac{1}{{\rm det} J}\right\vert_{z=z_j},\,\, {\rm
det} J \equiv \frac{\partial(x_1,x_2)}{\partial(y_1,y_2)}=1-{\partial\zeta\over \partial\bar{z}}
{\overline{\partial\zeta}\over \partial\bar{z}}.
\label{eqn:maggen}
\end{equation}
See {\em Witt} (1990) for a derivation of the rightmost equality. Note that these magnifications can be positive or negative, where the
sign corresponds to the parity (handedness) of the image. By definition in
microlensing the images are unresolved, and we are interested in the
total magnification which is just the sum of the magnifications of the
individual images, $A\equiv \sum_j |A_j|$.
An interesting and critical property of gravitational lenses is that
the mapping can be singular for some source positions. At these
source positions, ${\rm det} J = 0$. In other words, an
infinitesimally small displacement in the source position maps to an
infinitely large separation in the image position\footnote{This is analogous to the familiar
distortion seen in cylindrical
map projections of the globe (such as the Mercator projection) for
latitudes far from the equator, due to the singular nature of these
mappings at the poles.}. For point sources at these positions, the
magnification is formally infinite, and for sources near these
positions, the magnification is large.
From the lens equation (\ref{eqn:clenseq}),
\begin{equation}
{\partial\zeta\over \partial\bar{z}} = \sum_i^{N_l} \frac{\epsilon_i}{({\bar z}_{m,i} - {\bar z})^2},
\end{equation}
and therefore, from equation (\ref{eqn:maggen}), the image positions where ${\rm det} J = 0$ are given by,
\begin{equation}
\left| \sum_i^{N_l} \frac{\epsilon_i}{({\bar z}_{m,i} - {\bar z})^2} \right|^2 = 1.
\label{eqn:crit}
\end{equation}
The set of all such image positions define closed {\it critical curves}. These can be found by noting
that the sum in Equation \ref{eqn:crit} must have a modulus equal to unity. Therefore, we can solve
for the critical image positions parametrically by solving the equation
\begin{equation}
\sum_i^{N_l} \frac{\epsilon_i}{({\bar z}_{m,i} - {\bar z})^2} = e^{i\phi},
\label{eqn:critpar}
\end{equation}
for each value of the parameter $\phi = [0,2\pi)$.
The set of source positions corresponding to these image positions define closed curves called
{\it caustics}.
By clearing the complex conjugates ${\bar z}$ and fractions ({\em Witt}, 1990), Equation \ref{eqn:critpar} can
be written as a complex polynomial of degree $2N_l$, there are at most $2N_l$ critical curves and caustics.
Given their importance in planetary microlensing, caustics are discussed in considerably more
detail below.
\subsubsection{Single Lenses}\label{sec:point}
For a single point mass ($N_l=1$), defining the origin as the position of the
lens, the lens equation reduces to,
\begin{equation}
\beta = \theta - \frac{\theta_{\rm E}^2}{\theta}.
\label{eqn:singlec}
\end{equation}
Since the lens is circularly symmetric, the images are always located along the
line connecting the lens and source, and the vector notation has been suppressed, but
the convention that positive values of $\theta$ are for images on the same
side of the lens as the source is kept.
If the lens is perfectly aligned with the source, then $\beta=0$,
and $\theta=\theta_{\rm E}$. In other words, the lens images
the source into a ring of radius equal to the angular Einstein ring radius.
The dimensionless single lens equation is,
\begin{equation}
u=y-\frac{1}{y}.
\label{eqn:singles}
\end{equation}
In the case of imperfect alignment, this becomes a quadratic
function of $y$, and so there are two images of the source, with positions,
\begin{equation}
y_{\pm} = \pm \frac{1}{2}\left(\sqrt{u^2+4}\pm u\right).
\label{eqn:plimages}
\end{equation}
One of these images (the major image, or minimum) is always outside the Einstein ring radius ($y_+ \ge 1$)
on the same side of the lens as the source, and the other image (the minor image, or saddle point) is always inside the Einstein ring radius ($|y_-| \le 1$) on the opposite side of the lens as the source.
The separation between the two images is $|y_+-y_-|=(u^2+4)^{1/2}$, and thus the images
are separated by $\sim 2\theta_{\rm E}$ when both images are significantly magnified (i.e.\ when $u\la 1$).
Since $\theta_{\rm E}$ is of order a milliarcsecond for typical lens masses, and source and lens
distances in events toward the Galactic bulge, the images are unresolved.
The magnifications of each image can be found analytically,
\begin{equation}
A_\pm = \frac{1}{2}(A\pm 1)
\label{eqn:magind}
\end{equation}
where the total magnification is,
\begin{equation}
A(u)= \frac{u^2+2}{u\sqrt{u^2+4}}.
\label{eqn:magtot}
\end{equation}
A few properties are worth noting.
First, $u \ll 1$, $A(u)\simeq u^{-1}$. The magnification diverges for $u\rightarrow 0$, and the
point $u=0$ defines the caustic in the single lens case. Second, for $u \gg 1$, $A(u) \simeq 1+2u^{-4}$,
and thus the excess magnification drops rapidly for large source-lens angular separations.
\begin{figure}[h]
\epsscale{1.0}
\plotone{crit.eps}
\caption{\small
The critical values of $d$, the projected separation in units of $\theta_{\rm E}$, at which
the caustic topology (number of caustic curves) of a binary lens changes as a function of the mass ratio $q$.
The upper curve shows $d_w$, the critical value of $d$ between the wide caustic
topology consisting of two disjoint caustics, and the intermediate or resonant caustic topology consisting
of a single caustic. The lower curve shows $d_c$, the critical value between the resonant caustic
topology and the close caustic topology consisting of three disjoint caustics.
}\label{fig:crit}
\end{figure}
\subsubsection{Binary Lenses}\label{sec:binary}
For a two point-mass lens ($N_l=2$), the lens equation is,
\begin{equation}
\zeta = z + \frac{\epsilon_1}{{\bar z}_{m,1} - {\bar z}}+\frac{\epsilon_2}{{\bar z}_{m,2} - {\bar z}}.
\label{eqn:blenseq}
\end{equation}
This can be written as a fifth-order complex polynomial in $z$, which cannot
be solved analytically. {\em Witt \& Mao}, 1995 provide the coefficients of the
polynomial. It can easily be solved with standard numerical routines,
e.g., Laguerre's method ({\em Press et al.}, 1992), to yield the image
positions $z$. It is important to note that that the solutions
to the fifth-order complex polynomial are not necessarily solutions to the lens equation
(Equation \ref{eqn:blenseq}). Depending on the location of the source
with respect to the lens positions, two of the images can be spurious. Thus there
are either three or five images.
The boundaries of the three and five image regions are the caustic
curves (where ${\rm det} J =0$), and thus the number of images changes
by two when the source crosses the caustic. A binary lens has one,
two, or three closed and non-self-intersecting caustic curves. Which
of these three topologies is exhibited depends on the mass ratio of
the lens $q \equiv m_1/m_2$ and on the angular separation of the two
lens components in units of Einstein ring radius of binary, $d \equiv |z_{m,1}-z_{m,2}|$.
For a given $q$, the values of $d$ for which the topology changes are given by
({\em Schneider \& Weiss}, 1986; {\em Dominik}, 1999b),
\begin{equation}
\frac{q}{(1+q)^2} = \frac{(1-d_c)^3}{27d_c^8},\qquad d_w=\frac{(1+q^{1/3})^{3/2}}{(1+q)^{1/2}}.
\end{equation}
For $d\le d_c$, there are three caustic curves, for $d_c \le d \le d_w$, there
is one caustic curve, and for $d\ge d_w$, there are two caustic curves.
These are often referred to as the ``close'', ``intermediate'' or ``resonant'',
and ``wide'' topologies, respectively. Figure \ref{fig:crit} plots $d_c$ and $d_w$ as a function
of $q$. For equal-mass
binaries, $q=1$, the critical values of the separation are $d_c=2^{-1/2}$ and $d_w=2$.
A useful property of binary lenses is that the total magnification of all
the images is always $A\ge 3$ when the source is interior to the caustic
curve, i.e., when there are five images of the source ({\em Witt \& Mao}, 1995).
Thus if the magnification of a source is observed to be less than 3 during a microlensing
event, it can be immediately concluded that either the source is exterior to the caustic,
or the source is significantly blended with an unrelated, unlensed star\footnote{There
is a third, less likely possibility that there are additional bodies in the system. In
this case, the bound that $A\ge 3$ interior to the caustic can be violated ({\em Witt \& Mao}, 1995)}, thus diluting
the magnification.
\subsubsection{Triple Lenses and Beyond}\label{sec:triple}
For a triple lens ($N_l=3$), the lens equation can be written a
tenth-order complex polynomial in $z$ ({\em Gaudi et al.}, 1998), which can be solved using the
same techniques as in the binary lens case. {\em Rhie}, 2002 provides the coefficients of the
polynomial. There are a maximum of
ten images, and there are a minimum of four images, with the number
of images changing by a multiple of two when the source crosses the caustic.
The caustics of triple lenses can exhibit quite complicated topologies, including
nested and/or self-intersecting caustic curves. The topology depends
on five parameters: two mass ratios, two projected separations, and
the angle between projected position vectors of the companions and the
primary lens.
In general, the lens equation for a system of $N_l$ point lenses is
can be written as a complex polynomial of order $N_l^2+1$. However, it
has been shown that the maximum number of images is $5(N_l-1)$ for
$N_l\ge 2$ ({\em Rhie}, 2001, 2003; {\em Khavinson \& Neumann}, 2006). Thus for $N_l>3$, it is always
the case that some of the roots of the polynomial are not solutions to
the lens equation.
\subsubsection{Magnification Near Caustics}\label{sec:caustics}
As mentioned previously, the caustics are the set of source positions
for which the magnification of a point source is formally infinite.
Caustic curves are characterized by multiple concave segments called
{\it folds} which meet at points called {\it cusps}. Folds and cusps
are so named because the local lensing properties of sources close to
these caustics are equivalent to the generic fold and cusp
mapping singularities of mathematical catastrophe theory. For a more
precise formulation of this statement and additional discussion, see {\em Petters et al.}, 2001.
This leads to the particularly important property
of fold and cusps that their local lensing properties are universal, regardless
of the global properties of the lens. Thus the positions and magnifications
of the critical images of sources near folds and cusps have universal scaling behaviors
that can be described essentially analytically, and whose normalization
depend only on the local properties of the lens potential.
This universal and local behavior of the magnification
near caustics has proven quite useful, in that it allows one to analyze light curves
near caustic crossings separate from and independent of the global lens model (see,
e.g., {\em Gaudi \& Gould}, 1999, {\em Albrow et al.}, 1999b; {\em Rhie \& Bennett}, 1999, {\em Afonso et al.}, 2001;
{\em Dominik}, 2004a,b).
The lensing behavior near
folds and cusps has been discussed in detail by a number of authors ({\em Schneider \& Weiss}, 1986,1992;
{\em Mao}, 1992; {\em Zakharov}, 1995,1999;
{\em Fluke \& Webster}, 1999; {\em Petters et al.}, 2001, {\em Gaudi \& Petters}, 2002a,b;
{\em Pejcha \& Heyrovsky}, 2009). The salient properties are briefly reviewed
here, but the reader is encouraged to consult these papers for a more in-depth discussion.
For sources close to and interior\footnote{Here interior to a fold
caustic is defined such that the caustic curves away from the source.}
to a fold caustic, there are two highly-magnified images with nearly
equal magnification and opposite parity that are nearly equidistant
from the corresponding critical curve. Neglecting the curvature of
the caustic and any changes in the lensing properties parallel to the caustic,
the total magnification of these divergent images is ({\em Schneider \& Weiss}, 1986),
\begin{equation}
A_{\rm div}(\Delta u_\perp)= \left(\frac{ \Delta u_\perp}{u_r}\right)^{-1/2}\Theta(\Delta u_\perp),
\label{eqn:fold}
\end{equation}
where $\Delta u_\perp$ is the perpendicular distance to
the caustic, with $\Delta u_\perp>0$ for sources interior to the caustic, $\Theta(x)$ is the Heaviside step function,
and $u_r$ is the characteristic `strength' of the fold caustic locally, and is related to local
derivatives of the lens potential.
As the source approaches the caustic, the images
brighten and merge, disappearing when the source
crosses the caustic. The behaviors of the remaining (non-critical) images are
continuous as the source crosses the caustic. Thus immediately
outside of a fold, the magnification is finite and (typically) modest.
For sources close to and interior to a cusp, there are three
highly-magnified images. For a source on the axis of symmetry of the cusp, the
total magnification of the three images is $\propto \Delta u_c^{-1}$,
where $\Delta u_c$ is the distance of the source from the cusp. As the
source approaches the cusp, the magnification of all three images
increases and the images merge. Two of the three images disappear as the source exits
the cusp. The magnification of the remaining image is continuous
as the source exits the cusp, and in particular the image remains highly magnified, also with
magnification which is $\propto \Delta u_c^{-1}$. Thus, in contrast to
folds, sources immediately exterior to cusps are highly magnified.
As with the fold, the behaviors of the non-critical images are
continuous as the source crosses the caustic.
Caustic curves are closed, and thus for any given source trajectory
(which is simply a continuous path through the source plane), caustic
crossings come in pairs. Generally, since the majority of the length
of a caustic is made up of fold caustics, both the caustic entry and exit
are fold crossings. Since the magnification immediately outside of a
fold is not divergent, it is usually impossible to predict a caustic
entry beforehand. Once one sees a fold caustic entry, a caustic exit
is guaranteed, and this is typically a fold exit. Monitoring a
caustic exit is useful for two reasons. First, the strong finite
source effects (see below) during the crossing can be used to provide
additional information about the lens (see Section \ref{sec:properties}), and measure
the limb-darkening of the source (e.g., {\em Albrow et al.}, 1999c; {\em Fields et al.}, 2003).
In addition, during the caustic crossing the source is highly
magnified and potentially very bright, which allows for
otherwise impossible spectroscopic
observations to determine properties of the source star, such as
its effective temperature and atmospheric abundances ({\em Minniti et al.}, 1998;
{\em Johnson et al.}, 2008). Unfortunately, it is typically difficult to
predict when this exit will happen well before the crossing ({\em
Jaroszy{\'n}ski \& Mao}, 2001).
For both fold and cusp caustics, the magnification of a source of
finite size begins to deviate by more than a few percent from the
point-source approximation when the center of the source is
within several source radii of the caustic ({\em Pejcha \&
Heyrovsky}, 2009). Formally, the magnification of a finite source can
be found simply by integrating the point-source magnification over of
source, weighting by the source surface brightness distribution.
Practically, this approach is difficult and costly to implement
precisely due to the divergent magnification near the caustic curves.
An enormous amount of effort has been put into developing robust and
efficient algorithms to compute the magnification for finite sources
({\em Dominik}, 1995, 2007; {\em Wambsganss}, 1997; {\em Gould \&
Gaucherel}, 1997; {\em Griest \& Safizadeh}, 1998; {\em Vermaak},
2000; {\em Dong et al.}, 2006; {\em Gould}, 2008; {\em Pejcha \&
Heyrovsky}, 2009; {\em Bennett}, 2009b). The most efficient of these algorithms use a
two-pronged approach. First, semi-analytic approximations to the
finite-source magnification derived from an expansion of the
magnification in the vicinity of the source are used where appropriate
({\em Gould}, 2008; {\em Pejcha \& Heyrovsky}, 2009). Second, where
necessary a full numerical evaluation of the finite source
magnification is performed by integrating in the image plane.
Integrating in the image plane removes the difficulties with the
divergent behavior of the magnification near caustics, because the
surface brightness profiles of the images are smooth and continuous.
Thus one simply `shoots' rays in the image plane, and determines which
ones `land' on the source using the lens equation. The ratio of the
total area of all the images divided by the area of the source
(appropriately weighted by the surface brightness profile of the
source) gives the total magnification. The devilish details then lie
in the manner in which one efficiently samples the image plane
(see e.g., {\em Rattenbury et al.}, 2002; {\em Dong et al.}, 2006; {\em Pejcha \& Heyrovsky}, 2009; {\em Bennett}, 2009b).
For planetary microlensing events, efficient routines for evaluating
the finite-source magnification are crucial, for two reasons. First,
nearly all planetary perturbations are
strongly affected by finite source effects. Second, the processes of
finding the best-fit model to an observed light curve and evaluating
the model parameter uncertainties requires calculating tens of thousands of
trial model curves (or more), and the majority of the computation time is spent
calculating the finite-source magnifications.
\begin{figure*}[htp]
\epsscale{2.1}
\plotone{cfig3.ps}
\caption{\small
The grey curves show the caustics for a planetary lens with mass ratio $q=0.001$, and
various values of $d$, the projected separation in units of $\theta_{\rm E}$. The dotted
lines show sections of the Einstein ring. The dots show the location of the
planet. In panels c and i, an example trajectory is shown which produces a
perturbation by the planetary caustic; the resulting light curves are shown in Figure \ref{fig:plcurves}. In panel a,
three different representative angular source sizes in units of $\theta_{\rm E}$ are shown, $\rho_*=0.003, 0.01$,
and 0.03. For typical microlensing event parameters, these correspond to stars in the Galactic bulge with
radii of $\sim R_\odot, \sim 3R_\odot$, and $\sim 10R_\odot$, i.e., a main-sequence turn-off star, a subgiant, and a clump giant.
}\label{fig:pcaustics}
\end{figure*}
\subsection{Planetary Microlensing Phenomenology}\label{sec:phenom}
For binary lenses in which the companion mass ratio is $q\ll 1$ (i.e.,
planetary companions), the companion will cause a small perturbation to the
overall magnification structure of the lens. Thus the majority of
source positions will give rise to magnifications that are essentially
indistinguishable from a single lens. The source positions for which
the magnification deviates significantly from a single lens are all
generally confined to a relatively narrow region around the caustics.
Thus much of the phenomenology of planetary microlensing can be
understood by studying the structure of the caustics and the magnification
pattern near the caustics.
Recall there are three different caustic topologies for binary lenses:
close, intermediate, and wide (See Figure \ref{fig:crit}). For $q\ll 1$, the critical values of
the separation where these caustic topologies change can be
approximated by $d_c\simeq 1-3q^{1/3}/4$ and $d_w \simeq 1+3q^{1/3}/2$
({\em Dominik}, 1999b). Therefore, for planetary lenses, the
intermediate or resonant caustics are confined to a relatively
narrow range of separations near $d=1$, and this range shrinks as
$q^{1/3}$. Figure \ref{fig:pcaustics} shows the caustics for a Jupiter/Sun mass ratio
of $q=0.001$, and 11 different separations $d=0.6,0.7,0.8,0.9, 0.95,
1.0, 1.05, 1.11, 1.25, 1.43, 1.67$.
For both the close ($d<d_c$; Figure \ref{fig:pcaustics}i-k) and wide ($d>d_w$; Figure \ref{fig:pcaustics}a-c) topologies,
one caustic is always located near the position of the primary
(the origin in Figure \ref{fig:pcaustics}). This is known as the central caustic.
Figure \ref{fig:ccaustics} shows an expanded view of these caustics, which have a highly
asymmetric `arrow' shape, with one cusp at the arrow tip pointing
toward the planet, and three cusps at the `back end' of the caustic
pointing away from the planet. The on-axis cusp pointing away from
the planet is generally much `weaker' than the other cusps, in the
sense that the scale of the gradient in magnification along the cusp
axis is smaller for the weaker cusp, so that at fixed distance from
the cusp, the excess magnification is smaller for the weaker cusp.
The magnification pattern near a central caustic is illustrated in Figure \ref{fig:magmap}.
The light curves (one-dimensional slices through the magnification pattern)
for sources passing perpendicular to the binary-lens
axis close the back-end will exhibit a `U'-shaped double-peaked
deviation from the single-lens form (Figure \ref{fig:clcurves}c,d), whereas sources passing
perpendicular to the binary-lens axis close to the tip of the central
caustic will exhibit a single bump (Figure \ref{fig:clcurves}b,d). Sources passing the caustic
parallel to the binary-lens axis
will exhibit little deviation from the single lens form, essentially unless they
cross the caustic.
In the limit that $q\ll 1$, and $|d-1|\gg q$, it is possible show
that, for fixed $d$, the size of the central caustic scales as $q$.
Furthermore, the overall shape of the caustic, as quantified,
e.g., by its length-to-width ratio, depends only on $d$, such that the
caustic becomes more asymmetric (the length-to-width ratio increases)
as $d\rightarrow 1$. Finally, the central caustic shape and size is
invariant under the transformation $d\rightarrow d^{-1}$. See {\em
Chung et al.}, 2005 and Figure \ref{fig:ccaustics}. As illustrated
in Figure \ref{fig:clcurves}, the $d \rightarrow d^{-1}$ duality
results in a degeneracy between light curves produced by
central caustic perturbations due to planetary companions, typically referred to as the close/wide
degeneracy ({\em Griest \& Safizadeh} 1999; {\em Dominik}, 1999b; {\em Albrow et al.}, 2000;
{\em An}, 2005)
For the close $(d<d_c)$ topology (Figure \ref{fig:pcaustics}h-j), there are three caustics, the
central caustic and two larger, triangular-shaped caustics with three
cusps. The latter caustics are referred to as the planetary caustics,
and are centered on the planet/star axis at angular separation of $u_c
\simeq |d-d^{-1}|$ from the primary lens, on the opposite side of the
primary from the planet\footnote{It is possible to derive the location
of the planetary caustic(s) for both the close and wide topologies by noting
that, when the source crosses the planetary caustic(s), the planet is perturbing one of the two images
of the source created
by the primary lens (the major image in the case of the wide topology, and
the minor image in the case of the close topology).
For a source position $u$, the locations of the two primary
images $y_\pm (u)$ are given by Equation \ref{eqn:plimages}. The
planet must therefore be located at $d\sim y_{\pm}(u)$ to significantly perturb
the image, and thus the center of caustic $u_c$ is located
at the solution of the inversion of Equation
\ref{eqn:plimages}, i.e., $u(y_\pm=d)$, which yields $u_c=d-d^{-1}$.}.
The caustics are symmetrically displaced
perpendicular to the planet/star axis, with the separation between the
caustics increasing with decreasing $d$ and so increasing $u_c$. As
illustrated in Figure \ref{fig:magmap}, the magnification pattern near these caustics
is characterized by small regions surrounding the caustics where a
source exhibits a positive deviation from the single lens
magnification, and a large region between the caustics where a source
exhibits a negative deviation from the single lens magnification.
Figure \ref{fig:plcurves}c,d shows a representative light curve from a source passing
near the planetary caustics of a close planetary lens with $d=0.8$ and
$q=0.001$.
For the wide $(d>d_w)$ topology (Figure \ref{fig:pcaustics}h-j), there are two caustics, the
central caustic and a single planetary caustic with four cusps. As for the
close planetary caustics, the wide planetary caustic is centered on
the planet/star axis at an angular separation of $u_c \simeq
|d-d^{-1}|$ from the primary lens, but in this case on the same side
of the primary from the planet. The caustic is an asteroid shape,
with the length along the planet/star axis being generally longer than the width.
The asymmetry (i.e., length-to-width ratio) of the planetary caustic
increases as $d\rightarrow 1$. The magnification pattern near the
wide planetary caustic is characterized by large positive deviations
interior to the caustic, and lobes of positive deviation extending
outward along the axes of the four cusps, particularly along the
planet/star axis in the direction of the primary. There are
relatively small regions of slight negative deviation from the
single-lens magnification immediately outside the fold caustic between the
cusps. Figure \ref{fig:plcurves}a,b shows a representative light curve from a source
passing through the planetary caustic of a wide planetary lens with
$d=1.25$ and $q=0.001$.
\begin{figure}[htp]
\epsscale{1.0}
\plotone{cfig4.ps}
\caption{\small
The black curves show the central caustics for a planetary lens with $q=0.001$, and
various values of $d$, the projected separation in units of $\theta_{\rm E}$. The primary
lens is located at the origin, and so trajectories which probe
the central caustic correspond to events with small impact
parameter $u_0$, or events with high maximum magnification. The grey
curves show the central caustic for a mass ratio of $q=0.0005$, demonstrating
that the size of the central caustic scales as $q$. For $q\ll 1$, the central
caustic and proximate magnification patterns are essentially identical under the transformation $d \leftrightarrow d^{-1}$. The degree of asymmetry, i.e.\ the
length to width ratio, of the central caustic depends on $d$, such that the caustic becomes more asymmetric
as $d\rightarrow 1$. In panels c and d, example trajectories are shown which produce
perturbations by the central caustic; the resulting light curves are shown in Figure \ref{fig:clcurves}.
In panel a, a representative angular source size in units of $\theta_{\rm E}$ of $\rho_*=0.003$ is
shown. For typical microlensing event parameters, this correspond to a star in the Galactic bulge of
radius $\sim R_\odot$, i.e., a main-sequence turn-off star.
}\label{fig:ccaustics}
\end{figure}
\begin{figure*}
\epsscale{1.8}
\plotone{magmap.ps}
\caption{\small
The magnification pattern as a function of source position for a planetary companion with
$q=0.001$ and $d=1.25$ (top panel),
$d=1.0$ (middle panel), and $d=0.8$ (bottom panel), corresponding to wide, intermediate/resonant,
and close topologies, respectively. The greyscale shading denotes $2.5\log(1+\delta)$,
where $\delta$ is the fractional deviation from the single-lens (i.e., no planet) magnification. White
shading corresponds to regions with positive deviation from the single lens magnification,
whereas black shading corresponds to negative deviations. For the wide and close topology,
there are two regions of large deviations, corresponding to the central caustics located
at the position of the primary (the center of each panel), and the planetary caustics. For the intermediate/resonant
topology, there is only one large caustic, which produces relatively weak perturbations
for a large fraction of the caustic area.
}\label{fig:magmap}
\end{figure*}
In the wide $(d>d_w)$ and close $(d<d_c)$ cases, the planet is essentially perturbing one
of the two images created by the primary lens, and the other image (on
the other side of the Einstein ring) is essentially unaffected. In
this case, and in the limit that $q \rightarrow 0$, the lensing
behavior near the planet and the perturbed primary image is equivalent
to a single lens with pure external shear ({\em Gould \& Loeb}, 1992;
{\em Dominik}, 1999; {\em Gaudi \& Gould}, 1997a), i.e., a Chang-Refsdal lens ({\em Chang \&
Refsdal}, 1979). The Chang-Refsdal lens has been studied extensively (e.g.,
{\em Chang \& Refsdal}, 1984; {\em An}, 2005; {\em An \& Evans},
2006), and its properties are well-understood. The lens equation is,
\begin{equation}
\zeta = z - \frac{1}{\bar z}-\gamma {\bar z},
\label{eqn:crlenseq}
\end{equation}
where $\gamma$ is the shear and the origin is taken to be the location
of the planet. This is equivalent to a fourth-order
complex polynomial in $z$, which can be solved analytically, or numerically in the same manner as
the binary-lens case (see Section \ref{sec:theory}). There are 2 or 4 images, depending
on the source position. The correspondence between the Chang-Refsdal
lens and the wide/close planetary case is achieved by setting
$\gamma=d^{-2}$ and choosing the origin of the binary lens to be
$d-d^{-1}$ from the primary lens, and including in the Chang-Refsdal
approximation the magnification of the
unperturbed image created by the primary on the other side of the
Einstein ring. Note that $\gamma >1$ corresponds to the planetary
caustics for the close topology, whereas $\gamma<1$ corresponds
to the wide topology.
The properties of the caustics of a Chang-Refsdal lens (and thus the
planetary caustics of wide/close planetary lenses) can be studied
analytically. Based on expressions from {\em Bozza}, 2000, {\em Han},
2006 has studied the scaling of the planetary caustics. The overall
size of the wide $(d>1)$ planetary caustic is approximately $\propto
q^{1/2}d^{-2}$, and the length-to-width ratio is $\sim 1+d^{-2}$.
The shape is independent of $q$ in this approximation. The overall
size of the close $(d<1)$ planetary caustics are approximately
$\propto q^{1/2}d^{3}$, and their shape is also independent of
$q$. The vertical separation between the two planetary caustics in the
close topology is $\propto q^{1/2}d^{-1}$.
For the resonant case ($d_c\le d \le d_w$) there
exists a single, relatively large caustic with six cusps. For fixed
$q$, the resonant caustic is larger than either the central or planetary caustics.
The large size of these caustics results in a large
cross-section and thus an enhanced detection probability. Indeed, in the first
planet detected by microlensing, the source crossed a resonant caustic. The large
size also means that the light curve deviation can last a significant fraction
of the duration of the event.
However, resonant caustics are also `weak' in the sense that for a large fraction of the area
interior to or immediate outside the caustic,
the excess magnification relative to a single lens is small (see Figure \ref{fig:magmap}).
The exceptions to this are source positions in the vicinity the cusp located on the planet/star axis
pointing toward the planet, and source positions near the `back end' of the caustic
near the position of the primary, which are characterized by large negative deviations relative
to the single-lens magnification. The precise shape of the resonant caustic depends
sensitively on $d$, and thus
small changes in the value of $d$ lead to large changes in the caustic morphology, as can
be seen in Figure \ref{fig:pcaustics}. As a result, the effects of orbital motion, which result in a change
in $d$ over the course of the event, are expected to be more
important for resonant caustic perturbations. The size of resonant caustics scales
as $q^{1/3}$, in contrast to planetary caustics, which scale as $q^{1/2}$, and central
caustics, which scale as $q$.
\begin{figure}[htp]
\epsscale{1.0}
\plotone{lcurvesc.ps}
\caption{\small
Example light curves of planetary perturbations arising from
the source passing close to the central caustic in a high-magnification
event, for a planet/star mass ratio of $q=0.001$. Panel (a) shows
the overall light curve. The impact parameter of the event with respect to the
primary lens is $u_0=0.02$, corresponding to a peak magnification
of $A_{\rm max} \sim u_0^{-1}=50$. Panels (b-e) show zooms of the light curve peak. Two
different cases are shown,
one case of the wide planetary companion with $d=1.25$ (b,c), and a close planetary companion with $d=0.8$ (d,e).
These two cases satisfy $d \leftrightarrow d^{-1}$ and demonstrate the close/wide degeneracy.
The source passes close to the central
caustic; two example trajectories are shown in Figure \ref{fig:ccaustics} and the resulting
light curves including the planetary perturbations are shown in panels b-e.
The dotted line shows the magnification with no planet, whereas the
solid lines show the planetary perturbations with source
sizes of $\rho_*=0,0.003$, and $0.01$, (lightest to darkest).
In panel e, the light curve for $\rho_*=0.03$ is also shown. In this case,
the primary lens transits the source, resulting in a `smoothed' peak.
Although the planetary deviation is largely washed out, it is still detectable
with sufficiently precise photometry.
}\label{fig:clcurves}
\end{figure}
\begin{figure}[htp]
\epsscale{1.0}
\plotone{lcurvesp.ps}
\caption{\small
Example light curves of planetary perturbations arising from
the source passing close to the planetary caustic for a planet/star mass ratio of
$q=0.001$. Panels (a,c) show the overall light curves, whereas
panels (b,d) show zooms of the planetary deviation.
Two cases are shown,
one case of the wide planetary companion with $d=1.25$ (a,b), and a close planetary companion with $d=0.8$ (c,d).
In both cases, the impact parameter of the event with respect to the
primary lens is $u_0=0.3$.
The trajectories for the light curves displayed are shown in Figure \ref{fig:pcaustics}.
The dotted line shows the magnification with no planet, whereas the
solid lines show the planetary perturbations with source sizes
of $\rho_*=0, 0.003$, $0.01$, and $0.03$ (lightest to darkest).\vspace{25mm}
}\label{fig:plcurves}
\end{figure}
\section{PRACTICE OF MICROLENSING}\label{sec:practice}
\subsection{Light Curves and Fitting}\label{sec:lightcurves}
The apparent relative motion between the lens and the source gives
rise to a time-variable magnification of the source: a microlensing event.
Is is often (but not always, see Section \ref{sec:higher}) a good
approximation that the source, lens, and observer are in uniform,
rectilinear motion, in which case the angular separation between the
lens and source as a function of time can be written as,
\begin{equation}
u(t)= \left( \tau^2 +u_0^2\right)^{1/2},
\label{eqn:uoft}
\end{equation}
where $\tau \equiv (t-t_0)/t_{\rm E}$,
$t_0$ is the time of closest alignment, which is also the time of maximum magnification,
$u_0$ is the impact parameter of the event, and $t_{\rm E}$ is the Einstein ring crossing time,
\begin{equation}
t_{\rm E} \equiv \frac{\theta_{\rm E}}{\mu_{\rm rel}},
\end{equation}
where $\mu_{\rm rel}$ is the relative lens-source proper motion.
Figure \ref{fig:ped} shows the magnification as a function of time for
a microlensing event due to a single lens, with impact parameters of $u_0=0.01,0.1,0.2,...,0.9,1.0$, which serve to illustrate
the variety of light curve shapes.
Of course, what is observed is not the magnification, but the flux of a photometered
source as a function of time, which is given by,
\begin{equation}
F(t)= F_s A(t) + F_b.
\label{eqn:foft}
\end{equation}
Here $F_s$ is the flux of the microlensing star, and $F_b$ is the flux of any
unresolved light (or ``blended light'') that is not being lensed.
The latter can include light from a companion to the source, light from
unrelated nearby stars, light from a companion to the lens, and (most
interestingly) light from the lens itself. Microlensing experiments
are typically carried out toward crowded fields in order to maximize the event rate,
and therefore one often finds unrelated stars blended with the microlensed source
for typical ground-based resolutions of $\sim 1''$. Even in the most crowded
bulge fields, most unrelated background
stars are resolved at the resolution of the {\it Hubble Space Telescope (HST)}. Figure \ref{fig:blending}
shows the fields of two events as observed from the ground with typical seeing, with
{\it HST}, and with ground-based adaptive optics (AO).
\begin{figure}[htp]
\epsscale{0.56}
\plotone{gs.ps}
\caption{\small
Top panel: A 15''$\times$ 15'' $I$-band image of the field
of planetary microlensing event OGLE-2005-BLG-071 obtained the OGLE 1.3m Warsaw telescope
at Las Campanas Observatory in Chile. Second panel from top: Same
field as the top panel, but taken with the {\it Advanced Camera for Surveys}
instrument on the {\it Hubble Space Telescope} in the F814W filter.
Third panel from top: 22''$\times$ 22'' $H$-band image of the field
of planetary microlensing event MOA-2008-BLG-310 obtained with the CTIO/SMARTS2 1.3m telescope
at Cerro Tololo InterAmerican Observatory in Chile. Bottom panel:
Same field and filter as the panel above, but taken with the NACO instrument
on VLT. In all panels, the arrow indicates the microlensing target.
}\label{fig:blending}
\end{figure}
The observed flux as a function of time for a microlensing event due
to a single lens can be fit by five parameters: $t_0, u_0, t_{\rm E}$,
$F_s$, and $F_b$. It is important to note that several of these
parameters tend to be highly degenerate. There are only four gross
observable properties of a single-lens curve: $t_0$, the overall timescale of the event (i.e.,
$t_{FWHM}$), and the peak and baseline
fluxes. Thus $u_0, t_{\rm E}, F_s$, and $F_b$ tend to be highly
correlated, and are only differentiated by relatively subtle
differences between light curves with the same values of the gross
observables but different values of $u_0, t_{\rm E}, F_s$, and $F_b$ ({\em
Wozniak \& Paczynski}, 1997; {\em Han}, 1999; {\em Dominik}, 2009).
As a result of these degeneracies, when fitting to data it is often
useful to employ an alternate parametrization of the single-lens model
that is more directly tied to these gross observables, in order to avoid
strong covariances between the model parameters.
In practice, several different observatories using several different
filter bandpasses typically contribute data to any given observed
microlensing event. Since the flux of the source and blend will vary
depending the specific bandpass, and furthermore different
observatories may have different resolutions and thus different
amounts of blended light, one must allow for a different source and
blend flux for each filter/observatory combination. Thus the total
number of parameters for a generic model fit to an observed dataset is
$N_{nl}+2\times N_O$, where $N_{nl}$ is the number of (non-linear) parameters
required to specify the magnification as a function of time, and $N_O$
is the total number of independent datasets. Since the observed flux
is a linear function of $F_s$ and $F_b$, for a given set of $N_{nl}$
parameters which specify $A(t)$, the set of source and
blend fluxes can be found trivially using a linear least-squares fit.
Thus in searching for the best-fit model, one typically uses a hybrid
method in which the best-fit non-linear parameters (e.g., $t_0, u_0,
t_{\rm E}$ for a single lens) are varied using, e.g., a downhill-simplex,
Markov Chain Monte Carlo (MCMC), or grid-search method, and the specific best-fit
$F_s$ and $F_b$ values for each trial set of non-linear parameters are
determined via linear least squares.
The addition of a second lens component increases the number of model
parameters by (at least) three, and the complexity enormously. As
discussed in Section \ref{sec:phenom}, the magnification pattern (magnification as a
function of source position) for a binary lens is described two
parameters, $d$, the separation of lens components in units of the
Einstein ring of the lens, and $q$, the mass ratio of the lens. Four
additional parameters specify the trajectory of the source through the
magnification patters as a function of time: $t_0, u_0, t_{\rm E}$, and
$\alpha$. The first three are analogous to the single lens case:
$u_0$ is the impact parameter of the trajectory from the origin of the
lens in units of $\theta_{\rm E}$, $t_0$ is the time when $u=u_0$, and
$t_{\rm E}$ is the Einstein ring crossing time. Finally, $\alpha$
specifies the angle of the trajectory relative to the binary-lens
axis. Thus $6+2\times N_O$ parameters are required to specific the
light curve arising from a generic, static binary lens. There are
many possibles choices for the origin of the lens, as well as the mass
used for the normalization of $\theta_{\rm E}$. The optimal choice depends
on the particular properties of the lens being considered, i.e., wide
stellar binary, close stellar binary, or planetary system ({\em
Dominik}, 1999b). For planetary systems, one typically chooses the
location of the primary for the origin and normalizes $\theta_{\rm E}$ the
total mass of the system.
Light curves arising from binary lenses exhibit an astonishingly
diverse and complex phenomenology. While this diversity makes for a
rich field of study, it complicates the interpretation of observed
light curves mightily, for several reasons. First, other than a few
important exceptions (i.e., for planetary caustic perturbations, see Section \ref{sec:basic}), the salient features of binary and planetary
light curves have no direct relationship the canonical parameters of
the underlying model. Thus it is often difficult to choose initial
guesses for the fit parameters, and even if a trial solution is found,
it is difficult to be sure that all possible minima have been located.
Second, small changes in the values of the canonical parameters can
lead to dramatic changes the resulting light curve. In particular,
the sharp changes in the magnification that occur when the source
passes close to or crosses a caustic can make any goodness-of-fit
statistic such as $\chi^2$ very sensitive to small changes in the
underlying parameters. This, combined with the shear size of
parameter space, makes brute-force searches difficult
and time-consuming.
The complicated and highly corrugated shape of the $\chi^2$ surface
also causes many of the usual minimization routines (i.e.\ downhill
simplex) to fail to find the global or even local minimum. These
difficulties are compounded by the fact that the magnification of a
binary lens is non-analytic and time-consuming to calculate when
finite source effects are important.
Aside from their diverse phenomenology, binary lens light curves also
have the important property that they can be highly degenerate, in the
sense that two very different underlying lens models can produce very
similar light curves. These degeneracies can be accidental, in the
sense that that with relatively poor quality data or incomplete
coverage of the diagnostic light curve features, otherwise
distinguishable light curves can provide equally good statistical fits
to a given dataset ({\em Dominik \& Hirshfeld}, 1996; {\em Dominik},
1999; {\em Albrow et al.}, 1999b). This is particularly problematic
for caustic-crossing binary-lens light curves in which only one (i.e.,
the second) caustic crossing is observed. In this instance, it is
typically the case that very different models can fit a given
dataset\footnote{An important corollary is that, given an
observed first caustic crossing, it is very difficult to predict the
time of the second caustic crossing well in advance. See {\em Albrow
et al.}, 1999b and {\em Jaroszy{\'n}ski, \& Mao}, 2001 for further
discussion.}. The more
insidious degeneracies, however, are those that arise from
mathematical symmetries in the lens equation itself ({\em Dominik},
1999b). For example, in the limit of a very widely separated ($d_w\gg
1$) or very close ($d_c\ll 1$) binary lenses, Taylor expansion reveals
that the lens equations in the two cases
are identical (up to an overall coordinate translation) to order $d_c^2$ or $d_w^{-2}$ for $d_w \leftrightarrow
d_c^{-1}(1+q)^{1/2}$ ({\em Dominik}, 1999b; {\em Albrow et al.}, 2002;
{\em An}, 2005). Thus the magnifications are also identical to this order in these two cases.
Note that for $q_c,q_w \ll 1$, this degeneracy is
simply $d_c \leftrightarrow d_w^{-1}$. This degeneracy was discussed in the context
of planetary lenses in Section \ref{sec:phenom}. It is not known
if there exist analogous degeneracies for more complex (i.e., triple)
lenses. These mathematical degeneracies are insidious because, if the
model is deep within the limits where the degeneracies manifest
themselves, it is essentially impossible, even in principle, to
distinguish between the degenerate models with the photometric data
alone (but see {\em Gould \& Han}, 2000 for a way to resolve the
close/wide degeneracy with astrometry). Both classes of degeneracies
complicate the fitting of observed light curves. First, the existence
of these degeneracies implies that simply finding a fit does not imply
that the fit is unique. Second, these degeneracies may complicate the
physical interpretation of observed light curves, because the
degenerate models generally imply very different lens properties.
A number of authors have developed routines to locate fits to observed
binary lens light curves. {\em Mao \& DiStefano}, 1999 compiled a
large library of point-source binary lens light curves, and classified
these according to their salient features, such as, e.g., the number
of peaks and time between peaks. They then match the features in the
observed light curves to those in the library to find trial solutions
for minimization routines. This approach effectively works by
establishing a mapping between the light curve features and the
canonical parameters. {\em DiStefano \& Perna}, 1997 adopted a
conceptually similar approach, where they decomposed the observed
light curve into a linear combination of basis functions, then
compared the resulting coefficients of this fit to those found for
a library of events, again to identify promising trial solutions.
Similarly, {\em Vermaak}, 2003 used artificial neural networks to
identify promising regions of parameter space. While these methods
can in principle be applied to any binary-lens light curve, because they
are not intrinsically systematic, it is difficult to be sure that all possible fits
have been identified.
{\em Albrow et al.}, 1999b and {\em Cassan}, 2008 developed algorithms to
find a complete set of solutions for binary-lens light curves where the
source crosses a caustic. These algorithms are robust in the sense
that they will identify all possible fits to such light curves, and
thus uncover all possible degenerate solutions (e.g., {\em Afonso et
al.}, 2000). Unfortunately, they are also fairly user-intensive, and are
obviously only applicable to a limited subset of events.
Currently, the most robust and efficient approaches to fitting
observed binary-lens light curves use some variant of a hybrid approach,
for example a grid
search over those non-linear parameters that are not simply related to
the light curve features are thus are poorly-behaved (such as the mass
ratio $q$, projected separation $d$, and/or the angle of the trajectory
$\alpha$), combined with a downhill-simplex, steepest-descent, or MCMC fit to
the remaining non-linear parameters that are more directly related to
the observed light curve features. The linear parameters (such as the
source and blend flux) are trivially fit for each trial solution.
Often when there is some prior gross knowledge of the approximate
model parameters (i.e, a wide or close planetary or binary lens) a
judicious choice of parametrization guided by the inherent
magnification properties of the underlying lens geometry (i.e., the
caustic structure of the lens) can both speed up the fit and improve
the robustness of the fitting. Once trial solutions are identified,
they can be more carefully explored using, e.g., Markov Chain Monte Carlo techniques.
See, e.g., {\em Gould et al.}, 2006, {\em Dong et al.}, 2007, and {\em Bennett}, 2009.
Essentially all of these approaches must be
modified (although sometimes trivially) to include higher-order
effects when these provide significant perturbations to the observed
light curve (see Section \ref{sec:higher}).
In general, no robust, practical,
universal, and efficient algorithm exists for fitting an arbitrary
binary lens light curve in an automated way that is not highly
user-intensive. For higher-multiplicity (i.e., triple) lenses, the
algorithms are generally even less well-developed. As mentioned
in the introduction, this is likely the current single biggest impediment to the
progress of microlensing planet searches. Thus there is a
urgent and growing need for the development of (more automated) analysis software.
\subsection{Higher-Order Light Curve Effects}\label{sec:higher}
As reviewed in Section \ref{sec:lightcurves}, the simplest model of the observed light curve arising from
an isolated lens can be
described by $3+2\times N_O$ parameters: the three parameters that
describe the magnification as a function of time ($u_0,t_0, t_{\rm
E}$), and a blend and source flux for each of the $N_0$
observatory/filter combinations. This model, which accurately describes
the vast majority of observed microlensing light curves,
is derived under a number of assumptions, some of which may
break down under certain circumstances. When a light curve deviates from this
classic `Paczynski' or point-source single-lens form, it is generally classified as `anomalous'.
Not surprisingly, it is frequently the case that a robust measurement of these `anomalies' allows one to infer additional
properties of the primary, planet, or both (see Section \ref{sec:properties}).
The most obvious and most common anomaly is when the lens is not isolated.
Each additional lens component requires three additional parameters\footnote{If the additional component is very far away or very close to the primary lens,
the effects of the additional component can be described by only two additional parameters:
either the shear due the wide companion or the quadrupole moment of the lens for the close companion, and the angle of the trajectory
with respect to the projected binary axis.} (e.g., mass ratio, projected
separation, and angle between the additional component and the
primary), and so the light curve arising from a lens consisting of
$N_l$ masses requires a total of $3\times N_l+2\times N_O$ parameters.
Roughly $\sim 10\%$ of all microlensing lightcurves are observed to be due to binaries. In fact, the true fraction
of binary lenses is likely considerably higher, since the effects of a binary companion are only
apparent when $d$ is of order unity.
\noindent$\bullet$ {\bf Finite Source Effects}
The second most common anomaly occurs when the assumption of a point
source breaks down ({\em Gould}, 1994; {\em Witt \& Mao}, 1994; {\em
Witt}, 1995). As already discussed in Section \ref{sec:caustics}, this assumption
breaks down when the curvature of the magnification pattern is
significant over the angular size of the source, where `significant'
depends on the photometric precision with which the light curve is
measured. Generally, the center of the source must pass within a few
angular source radii $\theta_*$ of a caustic in order for finite
source effects to be important. The magnitude of finite source
effects are parametrized by the angular size of the source in units
of the angular Einstein ring radius, $\rho_*\equiv
\theta_*/\theta_{\rm E}$. For typical main-sequence sources in the
Galactic bulge, $\rho_*$ is of order $10^{-3}$, whereas for typical giant
sources, $\rho_*$ is of order $10^{-2}$. In addition, the
amount of (filter-dependent) limb-darkening can also have an important
effect on the precise shape of the portions of the light curve affected
by finite-source effects.
\noindent$\bullet$ {\bf Parallax}
Another common anomaly occurs when the assumptions of co-spatial
and/or non-accelerating observers break down. For example, the usual
expression for $u$, the angular separation between the lens and source
in units of $\theta_E$, presented in Equation \ref{eqn:uoft} assumes that the
observer (as well as the lens and source) are moving with constant
velocity. The assumption that the values of $t_0$ and $u_0$ are the
same for all observers implies that they are all co-spatial. When
either of these assumptions break down, this is generally termed
(microlens) parallax. This occurs in one of three ways.
First, when the duration of the microlensing event is a significant
fraction of a year, the acceleration of the Earth leads to a
significant non-uniform and/or non-rectilinear trajectory of the lens
relative to the source, which leads to deviations in the observed
light curve ({\em Gould}, 1992). This is referred to as `orbital
parallax'. Second, for sources very close to a caustic, small
parallax displacements due to the differences in the perspective leads
to differences in the magnifications seen by observers located at
different observatories for fixed time ({\em Holz \& Wald}, 1996).
This is commonly known as `terrestrial parallax'.
Finally, if the observers are separated by a significant fraction of an
AU, for example if the event is simultaneously observed from the Earth and a satellite in a solar orbit,
then the differences in the magnification are large even for moderate magnification ({\em Refsdal}, 1966;
{\em Gould}, 1994). This is known as `satellite parallax'.
In all three cases, the magnitude of the effect depends on the size of
relevant length scale that gives rise to the different or changing
perspective (i.e., the projected separation between the observatories,
the projected size of the Earth's orbit, or the projected separation between
the satellite and Earth) relative to ${\tilde
r}_{\rm E}\equiv D_{\rm rel}\theta_{\rm E}$, the angular Einstein ring radius projected on
the plane of the observer (See Figure \ref{fig:dia}).
\noindent$\bullet$ {\bf Xallarap}
Analogously to parallax, if the source undergoes significant
acceleration over the course of the event due to a binary companion,
this will give rise to deviations from the canonical light curve form
({\em Dominik}, 1998). This effect is commonly known as `xallarap',
i.e., parallax spelled backwards, to highlight the symmetry between
this effect and orbital parallax. In fact, it is always possible to
{\it exactly} mimic the effects of orbital parallax with a binary
source with the appropriate orbital parameters ({\em Smith et al.}
2003). For sufficiently precise data, such that the parallax/xallarap
parameters are well constrained, these two scenarios can be
distinguished because it would be {\it a priori} unlikely for a binary
source to have exactly the correct parameters to mimic the effects of
orbital parallax ({\em Poindexter et al.}, 2005). The magnitude of
xallarap effect depends on the semimajor axis of the binary
relative to ${\hat r}_{\rm E} \equiv D_s\theta_{\rm E}$, the Einstein ring radius projected on
the source plane.
\noindent$\bullet$ {\bf Orbital Motion}
For binary or higher multiplicity lenses, relative orbital motion of
the components can also give rise to deviations from the light curve
expected under the usual static lens assumption ({\em Dominik}, 1998).
For the general case of a binary lens and a Keplerian orbit, an
additional five parameters beyond the usual
static-lens parameters are needed to specify the light curve ({\em
Dominik}, 1998). Typically, however, only two additional parameters
can be measured: the two components of the relative projected velocity
of the lenses. These two components can be parametrized by the rate
of the change of the binary projected binary axis ${\dot d}$ and the
angular rotation rate of the binary $\omega$. The effect of the
latter is simply to rotate that magnification pattern of the lens on
the sky, whereas the former results in a change in the lens mapping
itself, and thus a change in the caustic structure and magnification
pattern.
\noindent$\bullet$ {\bf Binary Sources}
If the source is a binary and the lens happens to pass sufficiently
close to both sources, the light curve can exhibit a deviation from the
generic single-lens, single source form ({\em Griest \& Hu}, 1992).
For static binaries, the morphology of the deviation depends on the
flux ratio of the sources, their projected angular separation in units
of $\theta_{\rm E}$, and the impact parameter of the lens from the
source companion. By convention, such `binary source' effects are
conceptually distinguished from xallarap effects by the nature of the
deviation: if the deviation is caused by the source companion being
significantly magnified by the lens, is called a binary source effect,
if it caused by the acceleration due to the companion, it is called
xallarap. Of course, depending on the parameters of the source and
lens, light curves can exhibit only binary source effects, only
xallarap, or both xallarap and binary-source effects.
\noindent$\bullet$ {\bf Other Miscellaneous Effects}
A number of additional high-order effects have been discussed in the
literature, for example the effects of the finite physical size of the
lens ({\em Bromley}, 1996; {\em Agol}, 2002). In most cases,
these higher-order effects are expected to unobservable and/or extremely rare.
\subsection{Properties of the Detected Systems}\label{sec:properties}
For microlensing events due to single, isolated lenses, the parameters
that can routinely be measured are the time of maximum magnification
$t_0$, the impact parameter of the event in units of the Einstein ring
radius $u_0$, and the Einstein timescale $t_{\rm E}$, along with a source
flux $F_s$ and blend flux $F_b$ for each observatory/filter
combination. Of these parameters, $u_0$ and $t_0$ are simply
geometrical parameters and contain no physical information about the
lens. The Einstein timescale $t_{\rm E}$ is a degenerate combination of
the lens mass $M$, the relative lens-source parallax $\pi_{\rm rel}$,
and the relative lens-source proper motion $\mu_{\rm rel}$. Therefore
it is not possible to uniquely determine the mass and distance to the
lens from a measurement of $t_{\rm E}$ alone. The blend flux $F_b$
contains light from any source that is blended with the source,
including light from the lens if it is luminous. Unfortunately, for
the typical targets toward the Galactic bulge, ground-based images
typically contain light from other, unrelated sources, and it is not
possible to isolate the light from the lens (see Figure \ref{fig:blending}). Therefore, in the vast
majority of microlensing events, the mass, distance, and proper motion
of the lens are unknown.
As discussed below, for binary and planetary microlensing events it is
routinely possible to infer the mass ratio $q$, and $d$, the instantaneous
projected separation between the planet and star in units of
$\theta_{\rm E}$. However, the mass of the planet is typically not
known without a constraint on the primary mass. Furthermore, a
measurement of $d$ alone provides very little information about the
orbit, since $\theta_{\rm E}$ and the inclination, phase, and ellipticity
of the orbit are all unknown {\it a priori}.
For these reasons, when microlensing planet searches were first
initiated it was typically believed that detailed information about
individual systems would be very limited for planets detected in
microlensing light curves. This apparent deficiency was exacerbated by the
perception that the host stars would typically be
too distant and faint for follow-up observations.
Fortunately, in reality much more
information can typically be gleaned from a combination of a detailed
analysis of the light curve and follow-up, high-resolution
imaging, using the methods outlined below.
\noindent $\bullet${\bf Mass Ratio and Projected Separation}
First, the requirements for accurately measuring the minimum three
additional parameters needed to describe the light curves of binary
and planetary microlensing events are discussed. These
three additional parameters are the aforementioned $q$, $d$, and $\alpha$.
For planetary caustic perturbations, these parameters can essentially
be `read off' of the observed light curve. In this case, the gross
properties of the planetary perturbation can be characterized by three
observable quantities: the time of the planetary perturbation
$t_{0,p}$, the timescale of the planetary perturbation $t_{{\rm E},p}$, and the
magnitude of the perturbation $\delta_p$. These quantities then
simply and completely specify the underlying parameters $q,d,\alpha$,
up to a two-fold discrete degeneracy in $d$, corresponding to whether
the planet is perturbing the major or minor image, i.e, $d
\leftrightarrow d^{-1}$. Since these two situations result in very
different types of perturbations (see Figure \ref{fig:plcurves}), this discrete
degeneracy is easily resolved ({\em Gaudi \& Gould}, 1997b). If finite
source effects are important but not dominant, then there also exists
a continuous degeneracy between $q$ and $\rho_*$, stemming from the
fact that in this regime, both determine the width of the perturbation
$t_p$ ({\em Gaudi \& Gould}, 1997b). However, this degeneracy is
easily broken by good coverage and reasonably accurate photometry in
the wings of the perturbation. Finally, there is also a degeneracy
between major image ($d>1$) planetary caustic perturbations and a
certain class of binary-source events, namely those with extreme
flux ratio between the two sources. Specifically, it is always
possible to find a binary-source light curve that can exactly
reproduce the observables $t_{0,p}, t_{{\rm E},p}$, and $\delta_p$ ({\em Gaudi},
1998). This degeneracy is also easily broken by good coverage and
accurate photometry ({\em Gaudi}, 1998; {\em Gaudi \& Han}, 2004; {\em Beaulieu et al.}, 2006).
For central or resonant caustic perturbations in high-magnification events,
extracting the parameters $q,d,\alpha$ is typically more
complicated due to the fact that there is no simple, general relationship
between the salient features of the light curve perturbation and these
these parameters. Thus fitting these perturbations typically requires
a more sophisticated approach, as discussed in Section \ref{sec:lightcurves}. In addition,
there are a number of degeneracies that plague central caustic
perturbations. First, as discussed in \ref{sec:phenom} and illustrated in
Figures \ref{fig:ccaustics} and \ref{fig:clcurves}, there is a close/wide duality such that the central
caustic shape and associated magnification pattern are highly
degenerate under the transformation $d \leftrightarrow d^{-1}$ ({\em
Griest \& Safizadeh} 1999; {\em Dominik}, 1999b). This degeneracy
becomes more severe for very close/very wide planets, and in some
cases it is essentially impossible to distinguish between the two
solutions, even with extremely accurate photometry and dense coverage
of the perturbation (e.g., {\em Dong et al.}, 2009a). There also
exists a degeneracy between central caustic planetary perturbations,
and perturbations due to very close or very wide binary lenses. Very
close binaries have small, asteroid-shaped caustic located at the
center-of-mass of the system, whereas very wide binaries have small,
asteroid-shaped caustics near the positions of each of the lenses.
The gross features of central caustic perturbations can be reproduced
by a source passing by the asteroid-shape caustic produced by a
close/wide binary lens ({\em Dominik}, 1999, {\em Albrow et al.},
2000; {\em An}, 2005). Since close/wide binary lenses are themselves
degenerate under the transformation $d_w \leftrightarrow
d_c^{-1}(1+q)^{1/2}$, there is a four-fold degeneracy for
perturbations near the peak of high-magnification events. Fortunately,
the degeneracy between planetary central caustic perturbations and
close/wide binary lens perturbations can be resolved with good coverage
of the perturbation and accurate photometry ({\em Albrow et al.},
2000; {\em Han \& Gaudi}, 2008; {\em Han}, 2009).
\noindent $\bullet${\bf Einstein Ring Radius}
The requirement
for detecting a planet via microlensing is generally that the source must
pass reasonably close to the caustics produced by the planetary companion.
However, this is also basically the condition for finite source size effects to be important.
Thus for most planetary microlensing events, it is possible
to infer the angular size of the source in units of the angular Einstein
ring radius, $\rho_* \equiv \theta_*/\theta_{\rm E}$.
The angular size of the source can be estimated by its de-reddened
color and magnitude using empirical color-surface brightness relations
determined from angular size measurements of nearby stars ({\em van
Belle}, 1999; {\em Kervella et al.}, 1994). The source flux $F_s$ is
most easily determined by a fit to the microlensing light curve of the
form $F(t)=F_s A(t)+F_b$, as the variable magnification of the source
allows one to `deblend' the source and blend flux. Determining the
color of the source in this manner requires measurements in two
passbands, and thus while observations are typically focused on a
single passband (typically a far-red visible passband such as $R$ or
$I$), it is important to acquire a few points in a second filter to
determine the source color. The extinction toward the source can be
approximately determined by comparison to nearby red giant clump stars
({\em Yoo et al.}, 2004), which have a known and essentially constant
luminosity and intrinsic color (e.g., {\em Paczynski \& Stanek},
1998). An error in the extinction affects both the inferred color and
magnitude of the source, fortunately these have opposite and nearly
equal effects on the inferred value of $\theta_*$ ({\em Albrow et
al.}, 1999a).
Thus for events in which finite source effects are robustly
detected, it is possible to measure $\theta_{\rm E}$ ({\em Gould}, 1994).
This partially breaks the timescale degeneracy, since
\begin{equation}
\frac{M}{D_{\rm rel}}=\frac{c^2}{4G} \theta_{\rm E}^2.
\label{eqn:massdrel}
\end{equation}
The distance to the source is typically known approximately from
its color and magnitude (and furthermore the overwhelming majority of sources
are in the bulge), and so
a measurement of $\theta_{\rm E}$ essentially provides a mass-distance relation
for the lens.
\noindent $\bullet${\bf Light from the Lens}
Although the majority of the lenses that give rise to microlensing
events are distant and low-mass main-sequence stars, most are
nevertheless usually sufficiently bright that their flux can be
measured to relative precision of $\la 10\%$ with moderate-aperture (1-2m) telescopes and
reasonable ($10^2-10^4~{\rm s}$) exposure times, provided that the
light from the lens can be isolated. The fit to the
microlensing light curve gives the flux of the source $F_s$, and the
blend flux $F_b$. The latter contains the flux from any stars that
are not being lensed but are unresolved on the image, i.e., blended
with the source star. Generally, this blend flux can be decomposed
into several contributions,
\begin{equation}
F_b = F_l + F_{l,c} + F_{s,c} + \sum_j F_{u,j},
\label{eqn:fblend}
\end{equation}
where $F_l$ is the flux from the lens, $F_{l,c}$ is the flux from any
(blended) companions to the lens, $F_{s,c}$ is the flux from any
(blended) companions to the source, and $F_{u,j}$ is the flux from
each unrelated nearby star $j$ that is blended with the source. As
illustrated in Figure \ref{fig:blending}, at typical ground-based resolutions of $1''$
and in the extremely crowded target fields toward the bulge where
microlensing surveys are carried out, it is often the case that there
are several unrelated stars blended with the source star. Therefore,
the lens light cannot be uniquely identified based on such data
alone. At the higher resolutions of $0.05-0.1''$ available from
{\it HST} or ground-based AO
imaging, essentially all stars unrelated to the source our lens are
resolved. Since the source and lens must be aligned to $\la \theta_{\rm E}
\sim 1~{\rm mas}$ for a microlensing event to occur and the typical
relative lens-source proper motions are $\mu_{\rm rel} \sim 5-10~{\rm
mas~yr^{-1}}$ for microlensing events towards the bulge, the lens and
source will be blended in images taken within $\sim 10$ years of the
event, even at the resolution of {\it HST}. However, because the
microlensing fit gives $F_s$, the lens flux can be determined by
subtracting this flux from the combined unresolved lens+source flux in
the high-resolution image, assuming no blended companions to the lens
or source ({\em Bennett et al.}, 2007).
There are several potential complications to this procedure to
determine the lens flux. First, $F_s$ determined from the
microlensing fit will generally not be absolutely calibrated. Thus
the high-resolution photometric data must be `photometrically aligned'
to the microlens dataset. Typically this is done by matching stars
common to both sets of images. The accuracy of this alignment is
usually limited to $\sim 1\%$ due to the small number of common,
isolated stars available ({\em Dong et al.}, 2009b). It may also be the high-resolution images
are taken in a different filter than that for which the source flux $F_s$
is determined, necessitating a (model-dependent) color transformation
and introducing additional uncertainties (e.g., {\em Bennett et al.}, 2008).
Note that it is possible to
avoid this procedure entirely if the high-resolution images can be
taken at two different epochs with substantially different source star
magnifications. However, since this means at least one epoch must be
taken when the source is significantly magnified during the event,
this requires target-of-opportunity observations. Finally, any light
in excess of the source detected in the high-resolution images may be
attributed to close physical companions to the lens or source. In
some favorable cases (i.e., high magnification events), it is possible
to exclude these scenarios by the (lack of) second order effects the
companion would produce in the observed light curve. For example, a
binary companion to the lens produces a caustic which would
be detectable in sufficiently high-magnification events, whereas a
sufficiently close companion to the source would give rise to xallarap
effects which would be detectable in long timescale events ({\em Dong et al.}, 2009b).
A measurement of the flux of the lens in a single passband, along with
a model for extinction as a function of distance and a mass-luminosity
relationship, gives a mass-distance relationship for the lens ({\em Bennett et al.}, 2007). A
second measurement of the flux in a different passband can provide a
unique mass and distance to the lens, subject to the uncertainties in
the intrinsic color as a function of mass and the dust extinction
properties as a function of wavelength and distance.
\noindent $\bullet${\bf Proper Motion}
For typical values of $\mu_{\rm rel} \sim 5-10~{\rm mas/yr}$ for microlensing events toward the Galactic bulge,
after a few years, the lens and source will be displaced by $\sim 0.01$ arcseconds.
For luminous lenses, and using space telescope or AO imaging, it is possible
measure the relative lens-source proper motion, either by measuring
the elongation of the PSF or by measuring the difference in the
centroid in several filters if the lens and source have significantly
different colors ({\em Bennett et al.}, 2007). The proper motion can be combined
with the timescale to give the angular Einstein ring radius, $\theta_{\rm E}=\mu_{\rm rel} t_{\rm E}$.
\noindent $\bullet${\bf Microlens Parallax}
For some classes of events,
it is possible to obtain additional information about the lens
by measuring the microlensing parallax, $\mbox{\boldmath$\pi$}_{\rm E}$, a vector with
magnitude $|\mbox{\boldmath$\pi$}_{\rm E}| = \rm AU/\tilde r_{\rm E}$, and direction
of the relative lens-source proper motion. Recall
$\tilde r_{\rm E} \equiv D_{\rm rel}\theta_{\rm E}$ is the Einstein ring radius
projected onto the observer plane. As discussed
in Section \ref{sec:higher} microlens parallax effects arise in one of three varieties:
orbital parallax due to the acceleration of the Earth during the
event, terrestrial parallax in high-magnification events
observed by non-cospatial observers, and satellite parallax for events
observed from the ground and a satellite in solar orbit.
Orbital parallax deviations are generally only significant for events
with timescales that are a significant fraction of a year, and so long
as compared to the median timescale of $\sim 20~{\rm days}$.
Furthermore, the deviations due to orbital parallax are subject to an
array of degeneracies ({\em Gould et al.}, 1994; {\em Smith et al.},
2003; {\em Gould}, 2004), which can hamper the ability to extract
unique microlens parallax parameters. The severity of these
degeneracies depend on the particular parameters of the event in
question, but for most events it is the case that the only
robustly-measured effect in the light curve is an overall asymmetry,
which only yields one projection of $\mbox{\boldmath$\pi$}_{\rm E}$, namely
that in the direction perpendicular to the instantaneous Earth-Sun
acceleration vector at the time of the event ({\em Gould et al.}
1994). If the direction of the relative lens-source proper motion vector
$\mbox{\boldmath$\mu$}_{\rm rel}$
can be independently determined from the proper motion of a luminous lens, then it is possible
to determine the full $\mbox{\boldmath$\pi$}_{\rm E}$ vector,
since $\mbox{\boldmath$\pi$}_{\rm E}$ is parallel to $\mbox{\boldmath$\mu$}_{\rm rel}$.
Orbital parallax measurements made from two observatories are also
subject to several degeneracies, which have been studied by several
authors ({\em Refsdal}, 1966; {\em Gould}, 1994,1995; {\em Boutreux \&
Gould}, 1996; {\em Gaudi \& Gould}, 1997a). These can be resolved
in a number of ways ({\em Gould}, 1995; {\em Gould}, 1999; {\em Dong et al.}, 2007),
including observing from a third observatory, which allows one
to uniquely `triangulate' the parallax effects ({\em Gould}, 1994).
Similarly, terrestrial parallax measurements from only two observatories
are subject to degeneracies which can be resolved with simultaneous observations
from a third observatory that is not co-linear with the other two.
A measurement of the microlens parallax allows one to partially break
the timescale degeneracy and provides a mass-distance relation for the lens,
\begin{equation}
M D_{\rm rel}=\frac{c^2}{4G} \tilde
r_{\rm E}^2.
\label{eqn:drelmass}
\end{equation}
\noindent $\bullet${\bf Orbital Motion of the Planet}
In at least two
cases, the orbital motion of the planet during the microlensing event
has been detected ({\em Dong et al.}, 2009a; {\em Bennett et al.}, 2009). The effects of orbital motion generally allow the
measurement of the two components of the projected velocity of the
planet relative to the primary star. If an external measurement of the
mass of the lens is available, and under the assumption of a circular
orbit, these two components of the projected velocity completely
specify the full orbit of the planet (including inclination), up to a
two-fold degeneracy ({\em Dong et al.}, 2009a). In some cases, higher-order effects of
orbital motion can be used to break this degeneracy and even constrain
the ellipticity of the orbit ({\em Bennett et al.}, 2009).
\noindent $\bullet${\bf Bayesian Analysis}
In the cases when only $t_{\rm E}$, $q$, and $d$ can be measured, constraints on the
mass and distance to the lens (and so mass and semimajor axis of the planet),
must rely on a Bayesian analysis which incorporates priors on the distribution
of microlens masses, distances and velocities (e.g., {\em Dominik}, 2006; {\em Dong et al.}, 2006).
\noindent $\bullet${\bf Complete Solutions}
In many cases, several of these pieces of information can be measured
in the same event, often providing complete or even redundant measurements
of the mass, distance, and transverse velocity of the event. For example,
a measurement of $\theta_{\rm E}$ from finite source effects, when combined
with a measurement of $r_{\rm E}$ from microlens parallax, yields
the lens mass,
\begin{equation}
M=\left(\frac{c^2}{4G}\right)\tilde r_{\rm E} \theta_{\rm E},
\label{eqn:masssol}
\end{equation}
distance
\begin{equation}
D_{l}^{-1} = \frac{\theta_{\rm E}}{\tilde r_{\rm E}} + D_s^{-1},
\label{eqn:dlsol}
\end{equation}
and transverse velocity ({\em Gould}, 1996).
\subsection{Practical Aspects of Current Microlensing Searches}\label{sec:practical}
The microlensing event rate toward the Galactic bulge is ${\cal
O}(10^{-6})$ events per star per year ({\em Paczynski}, 1991). In a typical field toward the
Galactic bulge, the surface density of stars is $\sim 10^{7}$ stars
per deg$^2$ to $I\sim 20$ ({\em Holtzman et al.}, 1998). Thus to
detect $\sim 100$ events per year, $\sim 10$ deg$^2$ of the Bulge must
be monitored. Until relatively recently, large-format CCD cameras
typically had fields of view of $\sim 0.25$ deg$^2$, and thus $\sim
40$ pointings were required and so fields could only be monitored once
or twice per night. While this cadence is sufficient to detect the
primary microlensing events, it is insufficient to detect and
characterize planetary perturbations, which last a few days or less.
As a result, microlensing planet searches have operated using a
two-stage process. The Optical Gravitational Lens Experiment (OGLE,
{\em Udalski}, 2003) and the Microlensing Observations in Astrophysics
(MOA, {\em Sako et al.}, 2008) collaborations monitor several tens of
square degrees of the Galactic bulge, reducing their data real-time in
order to alert microlensing event in progress. A subset of these
alerted events are then monitored by several follow-up collaborations,
including the Probing Lensing Anomalies NETwork (PLANET, {\em Albrow
et al.}, 2008), RoboNet ({\em Tsapras et al.}, 2009), Microlensing
Network for the Detection of Small Terrestrial Exoplanets (MiNDSTEp,
{\em Dominik et al.}, 2008), and Microlensing Follow Up Network
($\mu$FUN, {\em Yoo et al.}, 2004) collaborations. Since only individual
microlensing events are monitored, these teams can achieve the sampling and
photometric accuracy necessary to detect planetary perturbations. In
fact, the line between the `alert' and `follow-up' collaborations is
now somewhat blurry, both because the MOA and OGLE collaborations
monitor some fields with sufficient cadence to detect planetary
perturbations, and because there is a high level of communication
between the collaborations, such that the observing strategies are
often altered real time based on available information about ongoing
events.
There are two conceptually different channels by which planets can be
detected with microlensing, corresponding to whether the planet is
detected through perturbations due to the central caustic (as shown in
Figure \ref{fig:clcurves}), or perturbation due to the planetary
caustic(s) (as shown in Figure \ref{fig:plcurves}). Because, for any
given planetary system, the planetary caustics are always larger than
the central caustics, the majority of planetary perturbations are
caused by planetary caustics. Thus searching for planets via the
influence of the planetary caustics is termed the `main channel'.
However, detecting planets via the main channel requires substantial
commitment of resources because the unpredictable nature of the
perturbation requires dense, continuous sampling, and furthermore the
detection probability per event is relatively low so many events must
be monitored. Detecting planets via their central caustic
perturbations requires monitoring high magnification primary events
near the peak of the event. In this case, the trade-off is that
although high-magnification events are rare (a fraction $\sim 1/A_{\rm
max}$ of events have maximum magnification $\ga A_{\rm max}$), they are individually
very sensitive to planets. The primary challenge lies with this
channel lies with identifying high-magnification events real time.
Which approach is taken depends on the resources that the individual
collaborations have available. The PLANET collaboration has
substantial access to 0.6-1.5m telescopes located in South Africa,
Perth, and Tasmania. With these resources, they are able to monitor
dozens of events per season, and so are able to search for planets via
the main channel. This tactic led to the detection of the first cool
rocky/icy exoplanet OGLE-2005-BLG-390Lb ({\em Beaulieu et al.}, 2006).
On the other hand, the $\mu$FUN collaboration use a single 1m
telescope in Chile to monitor promising alerted events in order try to
identify high-magnification events substantially before peak. When
likely high-magnification events are identified, the other telescopes in
the collaboration are then engaged to obtain continuous coverage of
the light curve during the high-magnification peak. High-magnification
events often reach peak magnitudes of $I \la 15$, and thus can
be monitored with relatively small-apertures (0.3-0.4m). This allows
amateur astronomers to contribute to the photometric follow-up. Indeed
over half of the members of the $\mu$FUN collaboration are amateurs.
\section{FEATURES OF THE MICROLENSING METHOD}\label{sec:features}
The unique way in which microlensing finds planets leads to some
useful features, as well as some (mostly surmountable) drawbacks.
Most of the features of the microlensing method can be understood
simply as a result of the fact that planet detection relies on the
direct perturbation of images by the gravitational field of the
planet, rather than on light from the planet, or the indirect effect
of the planet on the parent star.
\noindent$\bullet$ {\bf Peak Sensitivity Beyond the Snow Line}
The peak sensitivity of microlensing is for planet-star separations of
$\sim r_{\rm E}$, which corresponds to equilibrium temperatures of
\begin{equation}
T_{\rm eq} = 278~{\rm K} \left(\frac{L}{L_\odot}\right)^{1/4} \left(\frac{r_{\rm E}}{{\rm AU}}\right)^{-1/2}
\sim 70~{\rm K} \left(\frac{M}{0.5M_\odot}\right),
\label{eqn:Teq}
\end{equation}
where the rightmost expression assumes $L/L_\odot = (M/M_\odot)^5$, $D_l=4~{\rm kpc}$, and $D_s=8~{\rm kpc}$.
Thus microlensing is most sensitive to planets in the regions beyond the `snow line,' the
point in the protoplanetary disk
exterior to which the temperature is less than the condensation temperature of water
in a vacuum ({\em Lecar et al.}, 2006; {\em Kennedy et al.}, 2007; {\em Kennedy \& Kenyon}, 2008).
Giant planets are thought to form in the region immediately beyond the snow line, where the
surface density of solids is highest ({\rm Lissauer}, 1987).
Is microlensing sensitive to habitable planets? For assumptions above, and further assuming the habitable zone is centered
on $a_{HZ}={\rm AU}(L/L_\odot)^{1/2}$, the projected separation of a planet in
the habitable zone in units of $\theta_{\rm E}$ is $d_{HZ} \sim 0.25 (M/M_\odot)^2$. Thus for typical
hosts of $M\la M_\odot$, the habitable zone is well inside the Einstein ring radius ({\em Di Stefano}, 1999).
Since microlensing is much less sensitive to planets
with separations much smaller than the Einstein ring radius as these can only
perturb highly demagnified images, it is much less
sensitive to planets in the habitable zones of their parent stars ({\em Park et al.}, 2006)
for typical events.
However, it is important to note that this is primarily a statistical statement about
the typical sizes of the angular Einstein ring radii of microlensing events, rather
than a statement about any intrinsic limitations of the microlensing method.
A fraction of events have substantially smaller Einstein ring radii,
and for these events there is significant sensitivity to habitable planets.
In particular, lenses closer to the observer have smaller Einstein ring radii and so microlensing
events from nearby stars will have more sensitivity to habitable planets ({\em DiStefano \&
Night}, 2008; {\em Gaudi et al.} 2008b). Indeed,
a space-based microlensing planet search mission (described in Section \ref{sec:future}) will have significant
sensitivity to habitable planets, primarily due to the large number of events being
searched for planets ({\em Bennett et al.}, 2008).
\noindent$\bullet$ {\bf Sensitivity to Low-mass Planets}
The amplitudes of the perturbations caused by planets are typically large,
$\ga 10\%$, Furthermore,
although the durations of the perturbations get shorter with planet
mass (as $\sqrt{m_p}$)
and the probability of detection decreases (also roughly as $\sqrt{m_p}$),
the amplitude of the perturbations are independent of the planet mass.
This holds until the `zone of influence' of the planet, which has a
size $\sim \theta_{{\rm E},p}$, is smaller than the angular size of the source
$\theta_*$. When this happens, the perturbation is `smoothed'
over the source size, as demonstrated in Figures \ref{fig:clcurves} and \ref{fig:plcurves}. For typical parameters, $\theta_{{\rm E},p}\sim \mu{\rm
as}(m_p/M_\oplus)^{1/2}$, and for a star in the bulge,
$\theta_*\sim \mu{\rm as} (R_*/R_\odot)$.
This `finite source' suppression essentially precludes the detection
of planets with mass $\la 5~M_\oplus$ for clump giant sources in the bulge with $R_* \sim 10R_\odot$ ({\em Bennett \& Rhie}, 1996).
For main-sequence sources ($R\sim R_\odot$), finite source effects become
important for planets with the mass of the Earth, but does
not completely suppress the perturbations and render then undetectable until masses
of $\sim 0.02~M_\oplus \sim 2~M_{\rm Moon}$ for main-sequence sources ({\em Bennett \& Rhie}, 1996; {\em Han et al.}, 2005). Thus
microlensing is sensitive to Mars mass planets and even planets a few times the mass
of the Moon, for sufficiently small source sizes.
\noindent$\bullet$ {\bf Sensitivity to Long-Period and Free-Floating Planets}
Since microlensing can `instantaneously' detect planets without waiting
for a full orbital period, it is immediately sensitive to planets with very long
periods. Although the probability of detecting a planet decreases
for planets with separations larger than the Einstein ring radius
because the magnifications of the images decline, it does not drop to
zero. For events in which the primary star is also detected, the detection
probability for very wide planets with $d\gg 1$ is $\sim (\theta_{{\rm E},p}/\theta_{\rm E}) d^{-1} = q^{1/2} d^{-1}$
({\em Di Stefano \& Scalzo}, 1999b).
Indeed since microlensing is directly sensitive to the planet
mass, planets can be detected even without a primary microlensing event ({\em Di Stefano \& Scalzo}, 1999a).
Even free-floating planets that
are not bound to any host star are detectable in this way ({\em Han et al.}, 2005).
Microlensing is the only method that can detect old, free-floating
planets. A significant population of free-floating planets planets is a
generic prediction of most planet formation models, particular those
that invoke strong dynamical interactions to explain the observed
eccentricity distribution of planets ({\em Goldreich et al.}, 2004; {\em Juric \& Tremaine}, 2008;
{\em Ford \& Rasio}, 2004).
\begin{figure}[htp]
\epsscale{1.0}
\plotone{mass.eps}
\caption{\small
The rate of microlensing events toward the Galactic bulge as a function
of the mass of the lens, for main sequence (MS) stars and brown dwarfs (BD, $0.03M_\odot < M < 1 M_\odot$) (bold
dashed curve) and white dwarfs (WD), neutron stars (NS), and black hole (BH) remnants (solid curves). The total is shown by
a bold solid curve. From {\em Gould}, 2000.
}\label{fig:massdist}
\end{figure}
\noindent $\bullet$ {\bf Sensitivity to Planets Orbiting a Wide Range
of Host Stars}
The hosts probed by microlensing are
simply representative of the population of massive objects along the
line of sight to the bulge sources, weighted by the lensing
probability. Figure \ref{fig:massdist} shows a model by {\em Gould}, 2000 for the microlensing event
rate toward the Galactic bulge as a function of the host star
mass for all lenses, and also broken down by the type of
host, i.e., star, brown dwarf, or remnant. The sensitivity of microlensing is weakly
dependent on the host star mass, and has essentially no dependence on
the host star luminosity. Thus microlensing is about equally
sensitive to planets orbiting stars all along the main sequence, from
brown dwarfs to the main-sequence turn-off, as well as
planets orbiting white dwarfs, neutron stars, and black
holes.
\noindent $\bullet$ {\bf Sensitivity to Planets Throughout the Galaxy}
Because microlensing does not rely on light from the planet
or host star, planets can be detecting orbiting stars with distances
of several kiloparsecs. The microlensing event rate depends on the intrinsic
lensing probability, which peaks for lens distances about
halfway to the typical sources in the Galactic bulge, and the number density distribution of lenses
along the line-of-sight toward the bulge, which peaks at the bulge. The event rate remains substantial
for lens distances in the range $D_l\sim 1-8~{\rm kpc}$. Roughly
40\% and 60\% of microlensing events toward the bulge are expected to be
due to lenses in the disk and bulge, respectively ({\em Kiraga \& Paczynski}, 1994).
Specialized surveys may be sensitive to planets with $D_l \la 1~{\rm kpc}$
({\em Di Stefano}, 2008; {\em Gaudi et al.}, 2008),
as well as planets in M31 ({\em Covone et al.}, 2000; {\em Chung et al.}, 2006;
{\em Ingrosso et al.}, 2009).
\noindent $\bullet$ {\bf Sensitivity to Multiple-Planet Systems}
For low-magnification events, multiple planets in the same system
can be detected only if the source crosses the planetary caustics
of both planets, or equivalently only if both planets happen to have projected
positions sufficiently close to the paths of the two images created by
the primary lens. The probability of this is simply the product of the
individual probabilities, which is typically ${\cal{O}}(1\%)$ or less ({\em Han \& Park}, 2002). In
high-magnification events, however,
individual planets are detected with near-unity
probability regardless of the orientation of the planet with respect to
the source trajectory ({\em Griest \& Safizadeh}, 1998). This immediately implies all planets
sufficiently close to the Einstein ring radius will be revealed in
such events ({\em Gaudi et al.}, 1998). This, along with the fact that high-magnification events
are potentially sensitive to very low-mass planets, makes such events excellent probes
of planetary systems.
\noindent $\bullet$ {\bf Sensitivity to Moons of Exoplanets}
Because microlensing is potentially sensitive to planets with mass
as low as that of a few times the mass of the Moon ({\em Bennett \& Rhie}, 1996), it is
potentially sensitive to large moons of exoplanets. A system with a star of mass $M$, planet of mass $m_p$, and moon
of mass $m_m$ corresponds to a triple
lens with the hierarchy $m_m \ll m_p \ll M$. {\em Bennett \& Rhie}, 2002, {\em Han \& Han}, 2002,
{\em Han}, 2008, and {\em Liebig \& Wambsganss}, 2009 have all considered the detectability of moons of various masses,
orbiting planets of various masses. Generally, these studies find that massive terrestrial $(m_m \sim M_\oplus)$ moons (should
they exist) are readily detectable. Less massive moons, with masses similar to our Moon,
are considerably more difficult to detect, but may be detectable in next-generation microlensing planet
searches, particularly those from space, in some favorable circumstances (i.e., for sources
with small dimensionless source size $\rho_*$).
Analogous to planetary microlensing, moons with projected separation
much smaller than the Einstein radius of the planet $r_{{\rm E},p}\equiv \theta_{{\rm E},p}D_l$ are difficult to detect ({\em Han}, 2008).
Contrary to planetary microlensing, however, the signal of the moon does not
diminish for angular separations much greater than $r_{{\rm E},p}$ ({\em Han}, 2008),
although of course it becomes less likely that both the planetary and moon signal will be detected simultaneously.
A minimum requirement for a moon to be stable is that its semimajor axis
must be less the Hill radius of the planet, which is given by
\begin{equation}
r_H = a_p \left(\frac{m_p}{3M}\right)^{1/3},
\label{eqn:rhill}
\end{equation}
where $a_p$ is the semimajor axis of the planet. The ratio of $r_H$ to the Einstein ring
radius of the planet is,
\begin{equation}
\frac{r_H}{r_{{\rm E},p}} \sim d q^{-1/6}.
\label{eqn:rhillre}
\end{equation}
Thus for $q\la 10^{-3}$, detectable moons also (fortuitously) happen to be stable.
\section{RECENT HIGHLIGHTS}\label{sec:results}
\begin{figure*}[htp]
\epsscale{1.9}
\plottwo{ma.eps}{mt.eps}
\caption{\small
(Left) Mass versus semi-major axis for known exoplanets. (Right) Mass versus equilibrium temperature for known exoplanets.
In both panels, radial velocity detections are indicated by red circles, transiting planets
are indicated by blue triangles, microlensing detections are indicated
by green pentagons, and direct detections are indicated by magenta squares.
The letters indicate the locations of the Solar System planets. The solid pentagons are those microlensing detections for which the primary
mass and distance are measured, and thus planet mass, semimajor axis, and equilibrium
temperature are well-constrained. The open pentagons indicate those microlensing detections
for which only partial information about the primary properties is available,
and thus the planet properties are relatively uncertain. An extreme example
is shown in right panel, where two points connected
by a dotted line are plotted for event MOA-2008-BLG-310Lb ({\em Janczak et al.} 2009),
corresponding to the extrema of the allowed range of planet properties. In the right panel,
the light blue stripe roughly indicates the location of the
snow line. These figures also demonstrate the complementarity of the various planet detection
techniques: transit and radial velocity surveys are generally sensitive to planets
interior to the snow line, whereas microlensing and direct imaging surveys are sensitive to more distant
planets beyond the snow line.}
\label{fig:observed}
\end{figure*}
To date, ten detections of planets with microlensing have been
announced, in nine systems ({\em Bond et al.}, 2004; {\em Udalski et
al.}, 2005; {\em Beaulieu et al.}, 2006; {\em Gould et al.}, 2006;
{\em Gaudi et al.}, 2008a; {\em Bennett et al.}, 2008; {\em Dong et
al.}, 2009; {\em Janczak et al.}, 2009; {\em Sumi et al.}, 2009). In
addition, there are another seven events with clear, robust planetary
signatures that await complete analysis and/or publication. The
masses, separations, and equilibrium temperatures of the ten announced
planets are shown in Figure \ref{fig:observed}. We can expect to a
handful of detections per year at the current rate.
The right panel of this Figure \ref{fig:observed} demonstrates that
the first microlensing planet detections are probing a region of
parameter space that has not been previously explored by any method,
namely planets beyond snow line. As a result, although the total
number of planets found by microlensing to date is small in comparison
to the sample of planets revealed by the radial velocity and transit
methods, these discoveries have already provided important empirical
constraints on planet formation theories. In particular, the
detection of four cold, low mass ($5-20~M_\oplus$) planets amongst the
sample of microlensing detections indicates that these planets are
common ({\em Beaulieu et al.}, 2006; {\em Gould et al.}, 2006, {\em
Sumi et al.}, 2009). The detection of a Jupiter/Saturn analog also
suggests that solar system analogs are probably not rare ({\em Gaudi
et al.}, 2008). Finally, the detection of a low-mass planetary
companion to a brown-dwarf star suggests that such objects can form
planetary systems similar to those around solar-type main-sequence
stars ({\em Bennett et al.} 2008).
Another important lesson learned from these first few detections is
that it is possible to obtain substantially more information about the
planetary systems than previously thought. In all ten cases, finite source
effects have been detectable and so it has been possible to measure
$\theta_{\rm E}$, which yields a mass-distance relation for the
primary (see Section \ref{sec:properties}). Furthermore, for four systems, additional constraints allow
for a complete solution for the primary mass and distance, and so
planet mass.
\subsection{Individual Detections}
\label{sec:ind}
The first two planets found by
microlensing, OGLE-2003-BLG-235/MOA-2003-BLG-53Lb ({\em Bond et al.}, 2004), and
OGLE-2005-BLG-071Lb ({\em Udalski et al.}, 2005), are Jovian-mass objects with separations of $\sim
2-4~{\rm AU}$. While the masses and separations of these planets
are similar to many of the planets discovered via radial velocity
surveys, their host stars are generally less massive and so the
planets have substantially lower equilibrium temperatures of $\sim 50-70$~K,
similar to Saturn and Uranus.
The third and fourth planets discovered by microlensing are
significantly lower mass, and indeed inhabit a region of
parameter space that was previously unexplored by any
method. OGLE-2005-BLG-390Lb is a very low-mass planet with a
planet/star mass ratio of only $\sim 8 \times 10^{-5}$ ({\em Beaulieu et al.}, 2006). A Bayesian
analysis combined with a measurement of $\theta_{\rm E}$ from finite
source effects indicates that the planet likely orbits a low-mass M
dwarf with $M=0.22_{-0.11}^{+0.21}M_\odot$, and thus has a mass of
only $5.5_{-2.7}^{+5.5}M_\oplus$. Its separation is
$2.6_{-0.6}^{+1.5}~{\rm AU}$, and so has a cool equilibrium
temperature of $\sim 50~{\rm K}$. OGLE-2005-BLG-169Lb is another
low-mass planet with a mass ratio of $8\times 10^{-5}$ ({\em Gould et al.}, 2006), essentially
identical to that of OGLE-2005-BLG-390Lb. A Bayesian
analysis indicates a primary mass of $0.52_{-0.22}^{+0.19}M_\odot$,
and so a planet of mass $\sim 14_{-6}^{+5} M_\oplus$, a separation
of $3.3_{-0.9}^{+1.9}~{\rm AU}$, and an equilibrium temperature of
$\sim 70~{\rm K}$ ({\em Bennett et al.}, 2007). In terms of its mass and equilibrium temperature,
OGLE-2005-BLG-169Lb is very similar to Uranus. OGLE-2005-BLG-169Lb
was discovered in a high-magnification $A_{\rm max}\sim 800$ event; as argued
above, such events have significant sensitivity to multiple
planets. There is no
indication of any additional planetary perturbations in this event,
which excludes Jupiter-mass planets with separations between 0.5-15~AU,
and Saturn-mass planets with separations between 0.8-9.5~AU. Thus
it appears that this planetary system
is likely dominated by the detected Neptune-mass companion.
The fifth and sixth planets discovered by microlensing were detected
is what is arguably one of the most information-rich and complex microlensing
events ever analyzed, OGLE-2006-BLG-109 ({\em Gaudi et al.}, 2008; {\em Bennett et al.}, 2009). The light curve exhibited
five distinct features. Four of these features are attributable to the
source crossing the resonant caustic of a Saturn-mass planet, and
include a short cusp crossing, followed by a pair of caustic
crossings, and finally a cusp approach, all spanning roughly two weeks.
The fifth feature cannot be explained by the Saturn-mass planet
caustic, but rather is due to the source approaching the tip of the
central caustic of a inner, Jupiter-mass planet.
The OGLE-2006-BLG-109Lb planetary system bears a remarkable similarity
to a scaled version of Jupiter and Saturn. By combining all the
available information, it is possible to infer that the primary is an
M dwarf with a mass of $M=0.51^{+0.05}_{-0.04}~M_\odot$ at distance of
$D_l=1.51^{+0.11}_{-0.12}~{\rm kpc}$. The two planets have masses of $0.73\pm
0.06~M_{\rm Jup}$ and $0.27\pm 0.02~M_{\rm Jup}$ and separations of
$2.3 \pm 0.5~{\rm AU}$ and $4.5^{+2.1}_{-1.0}~{\rm AU}$, respectively. Thus
the mass and separation ratios of the two planets are very similar to
Jupiter and Saturn. Although the two planets in the
OGLE-2006-BLG-109Lb system orbit at about half the distance of their
Jupiter/Saturn analogs, their equilibrium temperatures are also similar
(although somewhat cooler) to Jupiter and Saturn, because of the lower
primary mass and luminosity.
The seventh planet detected by microlensing is also a very low-mass,
cool planet ({\em Bennett et al.}, 2008). MOA-2007-BLG-192Lb has a
mass ratio of $q= 2 \times 10^{-4}$, somewhat smaller than that of
Saturn and the sun. However, this event exhibits orbital parallax
and the best-fit model also shows evidence for finite source effects,
allowing a measurement of the primary mass. In this case, the
inferred primary mass is quite low:
$M=0.060_{-0.021}^{+0.028}~M_\odot$, implying a sub-stellar
brown-dwarf host, and very low mass for the planet of
$3.3_{-1.6}^{+4.9}~M_\oplus$. Remarkably, this planet was detected
from data taken entirely by the MOA and OGLE survey collaborations,
without the benefit of additional coverage from the follow-up
collaborations. This event therefore serves as an early indication of
how future microlensing planet surveys will operate, as detailed in
the next section. Unfortunately, in this case, coverage of the
planetary perturbation was sparse, and as a result the limits on the
mass ratio and primary mass are relatively weak. Fortunately, the inferred primary
mass can be confirmed and made more precise with follow-up VLT and/or
{\it HST} observations. If confirmed, this discovery will demonstrate that
brown dwarfs can form planetary systems similar to those around
solar-type main-sequence stars.
The eighth and ninth microlensing planets, MOA-2007-BLG-400Lb ({\em
Dong et al.}, 2008b) and MOA-2008-BLG-310Lb ({\em Janczak et al.},
2009) were detected in events that bear some remarkable similarities.
Both were high-magnification events ($A_{\rm max}\sim 628$ and $A_{\rm
max} \sim 400$, respectively) in which the primary lens transited the
source, resulting in a dramatic smoothing of the peak of the event
(see panel e of Figure \ref{fig:clcurves} for an example). By eye, single lens
models provide relatively good matches to the data. Nevertheless, weak
but broad and significant residuals to the single-lens model are
apparent in both cases, which are well-fit by perturbations due to the
central caustic of a planetary companion. In both cases, the inferred
caustic size is significantly smaller than the source size. Also in
both cases, precise constraints on the mass ratio are possible despite
the weak perturbation amplitudes. MOA-2007-BLG-400Lb is a cool,
Jovian-mass planet with mass ratio of $q=0.0025 \pm 0.0004$, and
MOA-2008-BLG-310Lb is sub-Saturn mass planet with $q=(3.3 \pm 0.3)
\times 10^{-4}$. The angular Einstein ring radius of the primary lens 2008-BLG-310-310L
as inferred from finite source effects is quite
small ($\theta_{\rm E} = 0.162 \pm 0.015~{\rm mas}$), implying that if
the primary is a star ($M\ge 0.08~M_\odot$), it must have $D_l > 6.0~{\rm kpc}$
and so be in the bulge ({\em Janczak et al.}, 2009). This is the best candidate yet for a bulge
planet. Unfortunately, it is not possible to definitely determine if the system is in the
bulge with currently-available information, and it is possible that
the primary is a low-mass star or brown dwarf in the foreground Galactic disk. This ambiguity
can be resolved with follow-up high resolution
imaging taken immediately, followed by additional observations taken in $\sim 10$ years, at which point the
source and lens will have separated by $\sim 50~{\rm mas}$.
The tenth planet, OGLE-2007-BLG-368Lb, is another cold
Neptune ({\em Sumi et al.,} 2009), making it the fourth
low-mass ($\la 20~M_\oplus$) planet discovered with microlensing. This
planet was detected in a relatively
low-magnification event with a maximum magnification of only $\sim 13$.
In this case, the source crossed between the two, triangular-shaped planetary caustics of
a close $(d<1)$ topology planetary lens, and thus the light curve exhibited a large ($\sim 20\%$) dip, characteristic
of the planetary perturbations due to such configurations (see Figure \ref{fig:plcurves}c,d).
\subsection{Frequency of Cold Planets}
\label{sec:freq}
The microlensing detection sensitivity declines with planet mass as
$\sim m_p^{1/2}$, and thus the presence of two low-mass planets
amongst the first four detections was an indication that the frequency
of cold super Earths/Neptunes ($5-15~M_\oplus$) is substantially
higher than that of cold Jovian-class planets. {\em Gould et al.}
2006 performed a quantitative analysis that accounted for the
detection sensitivities and Poisson statistics and demonstrated that,
at 90\% confidence, $38_{-22}^{+31}\%$ of stars host cold super
Earths/Neptunes Neptunes with separations in the range $1.6-4.3~{\rm
AU}$. An updated analysis taking into account the planets that have
since been detected, as well as those events that did not yield
planets, revises this number downward to $\sim 20\pm 10\%$ (D.\
Bennett, private communication). Thus, such planets are common, which
is ostensibly a confirmation of the core accretion model of planet
formation, which predicts that there should exist many more `failed
Jupiters' than bona-fide Jovian-mass planets at such separations,
particularly around low-mass primaries ({\em Laughlin et al.}, 2004;
{\em Ida \& Lin}, 2005).
By adopting a simple model for the scaling of the detection efficiency
with $q$, {\em Sumi et al.} 2009 used the distribution of mass ratios
of the ten microlensing planets discovered to date to derive an
intrinsic mass (ratio) function of exoplanets beyond the snow line of
${\rm d}N/{\rm d}q \propto q^{-1.7\pm 0.2}$. This implies that cold
Neptunes/super Earths ($q \sim 5 \times 10^{-5}$) are $7^{+6}_{-3}$
times more common than cold Jupiters ($q\sim 10^{-3}$), reinforcing
the conclusions of {\em Gould et el.} 2006.
\subsection{Properties of the Planetary Systems}
\label{sec:propobs}
As already mentioned, for all of the planet detections, it has been
possible to obtain additional information to improve the constraints
on the properties of the primaries and planets. For example, for
OGLE-2003-BLG-235/MOA-2003-BLG-53Lb, follow-up imaging with {\it HST}
yielded a detection of light from the lens, which constrains the mass
of the primary and planet to $\sim 15\%$,
$M=0.63_{-0.09}^{+0.07}M_\odot$ and $m_p=2.6_{-0.6}^{+0.8}~M_{\rm
Jup}$ ({\em Bennett et al.}, 2006). However, it is the analysis of
events OGLE-2006-BLG-109 ({\em Bennett et al.}, 2009) and
OGLE-2005-BLG-071 ({\em Dong et al.}, 2009a) which demonstrate most strikingly the detailed
information that can be obtained for
planets detected via microlensing.
For microlensing event OGLE-2006-BLG-109, besides the basic signatures of the two planets,
the light curve displays finite source effects and orbital
parallax, which allow for a complete solution to the lens system and
so measurements of the mass and distance to the primary, as well as a
the masses of the planets ({\em Gaudi et al.}, 2008a; {\em Bennett et al.}, 2009). This primary mass measurement is corroborated by a Keck
AO $H$-band image of the target that yields a measurement of the
flux of the lens, which in this case turns out to be brighter than the
source. The lens flux is consistent with the mass inferred entirely
from the microlensing observables. In addition, the orbital motion of
the outer Saturn-mass planet was detected. This was possible because the
source crossed or approached the resonant caustic due to this planet at four
distinct times. Furthermore, as discussed in Section \ref{sec:phenom},
the exact shape and size of resonant caustics depend sensitively on $d$,
which changes over the two weeks of the planetary deviations. In fact,
four of the six orbital parameters of the Saturn-mass planet are
well-measured, and a fifth parameter is weakly constrained. This information,
combined with an assumption of coplanarity with the Jupiter-mass planet and
the requirement of stability, enables a constraint on the orbital eccentricity of the Saturn-mass
planet of $e=0.15^{+0.17}_{-0.10}$, and inclination of the system of $i=64^{+4}_{-7}$ degrees
({\em Bennett et al.}, 2009).
In the case of OGLE-2005-BLG-071Lb, {\it HST} photometry, when combined with
information on finite source effects and
microlens parallax from the light curve, allows for a measurement of the primary
mass and distance, and so planet mass and projected separation. This
leads to the conclusion that the companion is a massive planet with
$m_p=3.8 \pm 0.4~M_{\rm Jup}$ and projected separation $r_\perp = 3.6 \pm 0.2~{\rm AU}$, orbiting
an M-dwarf primary with a mass of $M=0.46 \pm 0.04 M_\odot$ and a distance of
$D_l=3.2 \pm 0.4~{\rm kpc}$ ({\em Dong et al.}, 2009a). Furthermore, the primary
has thick-disk kinematics with a projected
velocity relative to the Local Standard of Rest of $v=103\pm 15~{\rm km~s^{-1}}$, suggesting that
it may be metal-poor. Thus, OGLE-2005-BLG-071Lb
may be a massive Jovian planet orbiting a metal-poor, thick-disk
M-dwarf. The existence of such a planet may pose a challenge for core-accretion models
of planet formation (e.g., {\em Laughlin et al.}, 2004; {\em Ida \& Lin}, 2004,2005).
Interestingly, all four microlensing planet hosts for which it has
been possible to measure the distance to the system lie in the
foreground disk. As mentioned above, MOA-2008-BLG-310Lb is the most
promising candidate for a bulge planet, but in this case the primary
could still be a foreground disk brown dwarf. In contrast, roughly
60\% of all microlensing events toward the Galactic bulge are due to
lenses in the bulge ({\em Kiraga \& Paczynski}, 1994). The lack of
confirmed microlensing bulge planets could be due to the selection effect
that longer events, which are more likely to arise from disk lenses, are
preferentially monitored by the follow-up collaborations, or it could
reflect a difference between the planet populations in the disk and
bulge. Regardless, with a larger sample of
planets with well-constrained distances, and a more careful accounting
of selection effects, it should be possible to compare the
demographics of planets in the Galactic disk and bulge.
\begin{figure*}
\epsscale{2.2}
\plottwo{ndetva.eps}{nvmass.eps}
\caption{\small
Expectations from the next generation of ground-based microlensing surveys,
including MOA-II, OGLE-IV, and a KMTNet telescope in South Africa.
These results represent the average of two independent simulations,
which include very different input assumptions
but differ in their predictions by only $\sim 0.3$ dex.
(Left) Number of planets detected per year as a function
of semimajor axis for various masses, assuming every star
has a planet of the given mass and semimajor axis.
(Right) Number of planets detected per year
as a function of planet mass, normalized by the number of $\sim 10M_\oplus$ found to date by microlensing
(indicated by the red dot). The solid line is the prediction assuming an equal number
of planets per logarithmic mass interval, and the dotted curve assumes
that the number of planets per log mass scales as $m_p^{-0.7}$ ({\em Sumi et al.}, 2009).
In both cases, planets are
assumed to be distributed
uniformly in $\log(a)$ between 0.4-20~AU. Arrows indicate the locations of
the ten published exoplanets.
}\label{fig:nextgen}
\end{figure*}
\section{FUTURE PROSPECTS}\label{sec:future}
In the four seasons of 2003-2006 there were five planetary events,
containing six detected planets. With the MOA upgrade in 2006 to the
MOA-II phase, the rate of planet detections has increased
substantially. From the 2007, 2008, and 2009 bulge seasons, there
were four, three, and four secure planetary events, respectively
(seven of these await publication). Thus even maintaining the current
rate, we can expect of order a dozen new planet detections over the
next several years. In fact, as described below, we can expect the rate of planet
detections to increase substantially, as microlensing planet searches
transition toward the next generation of surveys. Besides finding
more planets, these surveys will also have improved sensitivity to
lower-mass, terrestrial planets. Thus we can expect to have robust
constraints on the frequency of Earth-mass planets beyond the snow
line within the next decade. A space-based survey would determine
determine the demographics of planets with mass greater than that of Mars and
semimajor axis $\ga 0.5$AU, determine the frequency of free-floating,
Earth-mass planets, and determine the frequency of terrestrial planets in the
outer habitable zones of solar-type stars in the Galactic disk and
bulge.
The transition to the next generation
of ground-based surveys is enabled by the advent of large-format cameras with fields-of-view (FOV) of several
square degrees. With such large FOVs, it becomes possible to monitor tens of
millions of stars every $10-20$ minutes, and so discover thousands of
microlensing events per year. Furthermore, these events are then
simultaneously monitored with the cadence required to detect
perturbations due to very low-mass ($\sim M_\oplus$) planets. Thus
next-generation searches will operate in a very different mode than
the current alert/follow-up model. In order to obtain round-the-clock
coverage and so catch all of the perturbations, several such
telescopes would be needed, located on 3-4 continents roughly evenly
spread in longitude.
In fact, the transition to the `next-generation' is happening already.
The MOA-II telescope in New Zealand (1.8m and 2 deg$^2$ FOV)
already represents one leg of such a survey. The OGLE team has recently
upgraded to the OGLE-IV phase with a 1.4 deg$^2$ camera, which will represent the second
leg in Chile when it becomes operational in 2010. Although the OGLE telescope
has a smaller FOV camera and a smaller aperture (1.3m) than the
MOA telescope, these are mostly compensated by the
better site quality in Chile. Finally, the
Korean Microlensing Telescope Network (KMTNet) is
a project with plans to build three 1.6m telescopes with 4 deg$^2$ FOV
cameras, one each in South Africa, South America, and Australia. With
the completion of the South African leg of this network (planned
2012), a next-generation survey would effectively be in place. In addition, astronomers from Germany and China
are considering initiatives to secure funding to build 1-2m class
telescopes with wide FOV cameras in southern Africa or Antarctica.
Detailed simulations of such a next-generation microlensing survey
have been performed by several groups ({\em Gaudi et al.,
unpublished}; {\em Bennett}, 2004). These simulations include models
for the Galactic population of lenses and sources that match all
constraints ({\em Han \& Gould}, 1995, 2003), and account for
real-world effects such as weather, variable seeing, moon and sky
background, and crowded fields. They reach similar conclusions. Such
a survey would increase the planet detection rate at fixed mass by at
least an order of magnitude over current surveys. Figure
\ref{fig:nextgen} shows the predictions of these simulations for the
detection rate of planets of various masses and separations using a
survey including MOA-II, OGLE-IV, and a Korean telescope in South
Africa. In particular, if Earth-mass planets with semimajor axes of
several AU are common around main-sequence stars, a next generation
microlensing survey should detect several such planets per year. This
survey would also be sensitive to free-floating planets, and would
detect them at a rate of hundreds per year if every star has ejected
Jupiter-mass planet.
Ultimately, however, the true potential of microlensing cannot be
realized from the ground. Weather, seeing, crowded fields, and
systematic errors all conspire to make the detection of
planets with mass less than Earth effectively impossible from the
ground ({\em Bennett}, 2004). As outlined
in {\em Bennett \& Rhie}, 2002, a space-based microlensing survey offers several advantages:
the main-sequence bulge sources needed to detect sub-Earth
mass planets are resolved from space, the events can be monitored
continuously, and it is possible to observe the moderately reddened
source stars in the near infrared to improve the photon collection
rate. Furthermore, the high spatial resolution afforded by space
allows unambiguous identification of light from the primary
(lens) stars and so measurements of the primary and planet
masses ({\em Bennett et al.}, 2007).
The expectations from a Discovery-class space-based microlensing
survey are impressive. Such a survey would be sensitive to all
planets with mass $\ga 0.1M_\oplus$ and separations $a \ga 0.5~{\rm
AU}$, including free-floating planets ({\em Bennett \& Rhie}, 2002;
{\em Bennett et al.}, 2009). This range includes analogs to all the
solar system planets except Mercury. If every main-sequence star has
an Earth-mass planet in the range $1-2.5~{\rm AU}$, the survey would
detect $\sim 500$ such planets within its mission lifetime. The
survey would also detect a comparable number of habitable Earth-mass
planets as the {\it Kepler} mission ({\em Borucki et al.}, 2003), and so would provide an important
independent measurement of $\eta_{\oplus}$. When combined with
complementary surveys (such as {\it Kepler}), a space-based
microlensing planet survey would determine
the demographics of both bound and free-floating planets with masses greater than that
of Mars orbiting stars with masses less than that of the Sun.
The basic requirements for a space-based microlensing planet survey
are relatively modest: an aperture of at least 1m, a large FOV camera
(at least $\sim$0.5 deg$^2$) with optical or near-IR detectors,
reasonable image quality of better than $\sim 0.25$'', and an orbit
for which the Galactic bulge is continuously visible. The {\it
Microlensing Planet Finder (MPF)} is an example of a space-based
microlensing survey that can accomplish these objectives, essentially
entirely with proven technology, and at a cost of $\sim$\$300 million
excluding launch vehicle ({\em Bennett et al.}, 2009).
Interestingly, the requirements for a space-based microlensing planet
survey are very similar to, or less stringent than, the requirements
for a number of the proposed dark energy missions, in particular those
that focus on weak lensing measurements. Thus, it may be attractive
to consider a combined dark energy/planet finding mission that could
be accomplished at a substantial savings compared to doing each
mission separately ({\em Gould}, 2009).
\bigskip
\textbf{ Acknowledgments.} I would like to thank Subo Dong for preparing Figure \ref{fig:blending},
Andrzej Udalski for reading over the
manuscript, and two anonymous referees for catching some important
errors, and comments that greatly improved the presentation.
\bigskip
\centerline\textbf{ REFERENCES}
\bigskip
\parskip=0pt
{\small
\baselineskip=11pt
\par\noindent\hangindent=1pc\hangafter=1 Afonso, C., Alard, C., Albert, J.~N., Andersen, J., Ansari, R.,
Aubourg, {\'E}., Bareyre, P., Bauer, F., Beaulieu, J.~P., Bouquet, A.,
Char, S., Charlot, X., Couchot, F., Coutures, C., Derue, F., Ferlet, R.,
Glicenstein, J.~F., Goldman, B., Gould, A., Graff, D., Gros, M.,
Haissinski, J., Hamilton, J.~C., Hardin, D., de Kat, J., Kim, A., Lasserre,
T., Lesquoy, {\'E}., Loup, C., Magneville, C., Marquette, J.~B., Maurice,
{\'E}., Milsztajn, A., Moniez, M., Palanque-Delabrouille, N., Perdereau,
O., Pr{\'e}vot, L., Regnault, N., Rich, J., Spiro, M., Vidal-Madjar, A.,
Vigroux, L., Zylberajch, S., Alcock, C., Allsman, R.~A., Alves, D.,
Axelrod, T.~S., Becker, A.~C., Cook, K.~H., Drake, A.~J., Freeman, K.~C.,
Griest, K., King, L.~J., Lehner, M.~J., Marshall, S.~L., Minniti, D.,
Peterson, B.~A., Pratt, M.~R., Quinn, P.~J., Rodgers, A.~W., Stetson,
P.~B., Stubbs, C.~W., Sutherland, W., Tomaney, A., Vandehei, T., Rhie,
S.~H., Bennett, D.~P., Fragile, P.~C., Johnson, B.~R., Quinn, J., Udalski,
A., Kubiak, M., Szyma{\'n}ski, M., Pietrzy{\'n}ski, G., Wo{\'z}niak, P.,
Zebru{\'n}, K., Albrow, M.~D., Caldwell, J.~A.~R., DePoy, D.~L., Dominik,
M., Gaudi, B.~S., Greenhill, J., Hill, K., Kane, S., Martin, R., Menzies,
J., Naber, R.~M., Pogge, R.~W., Pollard, K.~R., Sackett, P.~D., Sahu,
K.~C., Vermaak, P., Watson, R.,
\& Williams, A.\ (2000) Combined Analysis of the Binary Lens Caustic-crossing Event MACHO 98-SMC-1. {\em ApJ, 532}, 340-352
\par\noindent\hangindent=1pc\hangafter=1 Afonso, C., Albert, J.~N., Andersen, J., Ansari, R., Aubourg, {\'E}.,
Bareyre, P., Bauer, F., Blanc, G., Bouquet, A., Char, S., Charlot, X.,
Couchot, F., Coutures, C., Derue, F., Ferlet, R., Fouqu{\'e}, P.,
Glicenstein, J.~F., Goldman, B., Gould, A., Graff, D., Gros, M.,
Haissinski, J., Hamilton, J.~C., Hardin, D., de Kat, J., Kim, A., Lasserre,
T., LeGuillou, L., Lesquoy, {\'E}., Loup, C., Magneville, C., Mansoux, B.,
Marquette, J.~B., Maurice, {\'E}., Milsztajn, A., Moniez, M.,
Palanque-Delabrouille, N., Perdereau, O., Pr{\'e}vot, L., Regnault, N.,
Rich, J., Spiro, M., Vidal-Madjar, A., Vigroux, L., Zylberajch, S.,
\& The EROS collaboration (2001) Photometric constraints on microlens spectroscopy of EROS-BLG-2000-5. {\em A\&A, 378}, 1014-1023
\par\noindent\hangindent=1pc\hangafter=1 Albrow, M., Beaulieu, J.-P., Birch, P., Caldwell, J.~A.~R., Kane, S.,
Martin, R., Menzies, J., Naber, R.~M., Pel, J.-W., Pollard, K., Sackett,
P.~D., Sahu, K.~C., Vreeswijk, P., Williams, A., Zwaan, M.~A.,
\& The PLANET Collaboration (1998) The 1995 Pilot Campaign of PLANET: Searching for Microlensing Anomalies through Precise, Rapid, Round-the-Clock Monitoring. {\em ApJ, 509}, 687-702
\par\noindent\hangindent=1pc\hangafter=1 Albrow, M.~D., Beaulieu, J.-P., Caldwell, J.~A.~R., Depoy, D.~L.,
Dominik, M., Gaudi, B.~S., Gould, A., Greenhill, J., Hill, K., Kane, S.,
Martin, R., Menzies, J., Naber, R.~M., Pollard, K.~R., Sackett, P.~D.,
Sahu, K.~C., Vermaak, P., Watson, R., Williams, A.,
\& Pogge, R.~W.\ (1999a) The Relative Lens-Source Proper Motion in MACHO 98-SMC-1. {\em ApJ, 512}, 672-677
\par\noindent\hangindent=1pc\hangafter=1 Albrow, M.~D., Beaulieu, J.-P., Caldwell, J.~A.~R., Depoy, D.~L.,
Dominik, M., Gaudi, B.~S., Gould, A., Greenhill, J., Hill, K., Kane, S.,
Martin, R., Menzies, J., Naber, R.~M., Pogge, R.~W., Pollard, K.~R.,
Sackett, P.~D., Sahu, K.~C., Vermaak, P., Watson, R., Williams, A.,
\& The PLANET Collaboration (1999b) A Complete Set of Solutions for Caustic Crossing Binary Microlensing Events. {\em ApJ, 522}, 1022-1036
\par\noindent\hangindent=1pc\hangafter=1 Albrow, M.~D., Beaulieu, J.-P., Caldwell, J.~A.~R., Dominik, M.,
Greenhill, J., Hill, K., Kane, S., Martin, R., Menzies, J., Naber, R.~M.,
Pel, J.-W., Pollard, K., Sackett, P.~D., Sahu, K.~C., Vermaak, P., Watson,
R., Williams, A.,
\& Sahu, M.~S.\ (1999c) Limb Darkening of a K Giant in the Galactic Bulge: PLANET Photometry of MACHO 97-BLG-28. {\em ApJ, 522}, 1011-1021
\par\noindent\hangindent=1pc\hangafter=1 Albrow, M.~D., An, J., Beaulieu, J.-P., Caldwell, J.~A.~R., DePoy,
D.~L., Dominik, M., Gaudi, B.~S., Gould, A., Greenhill, J., Hill, K., Kane,
S., Martin, R., Menzies, J., Pogge, R.~W., Pollard, K.~R., Sackett, P.~D.,
Sahu, K.~C., Vermaak, P., Watson, R.,
\& Williams, A.\ (2002) A Short, Nonplanetary, Microlensing Anomaly: Observations and Light-Curve Analysis of MACHO 99-BLG-47. {\em ApJ, 572}, 1031-1040
\par\noindent\hangindent=1pc\hangafter=1 Alcock, C., Akerlof, C.~W., Allsman, R.~A., Axelrod, T.~S., Bennett,
D.~P., Chan, S., Cook, C.~H., Freeman, K.~C., Griest, K., Marshall, S.~L.,
Park, H.~S., Perlmutter, S., Peterson, B.~A., Pratt, M.~R., Quinn, P.~J.,
Rodgers, A.~W., Stubbs, C.~W.,
\& Sutherland, W.\ (1993) Possible Gravitational Microlensing of a Star in the Large Magellanic Cloud. {\em Nature, 365}, 621-623
\par\noindent\hangindent=1pc\hangafter=1 Alcock, C., Allsman, R.~A., Alves, D., Axelrod, T.~S., Becker, A.~C.,
Bennett, D.~P., Cook, K.~H., Freeman, K.~C., Griest, K., Guern, J., Lehner,
M.~J., Marshall, S.~L., Peterson, B.~A., Pratt, M.~R., Quinn, P.~J., Reiss,
D., Rodgers, A.~W., Stubbs, C.~W., Sutherland, W., Welch, D.~L.,
\& The MACHO Collaboration (1996) Real-Time Detection and Multisite Observations of Gravitational Microlensing. {\em ApJ, 463}, L67-L70
\par\noindent\hangindent=1pc\hangafter=1 Alcock, C., Allsman, R.~A., Alves, D., Axelrod, T.~S., Becker, A.~C.,
Bennett, D.~P., Cook, K.~H., Freeman, K.~C., Griest, K., Keane, M.~J.,
Lehner, M.~J., Marshall, S.~L., Minniti, D., Peterson, B.~A., Pratt, M.~R.,
Quinn, P.~J., Rodgers, A.~W., Stubbs, C.~W., Sutherland, W., Tomaney,
A.~B., Vandehei, T.,
\& Welch, D.\ (1997) First Detection of a Gravitational Microlensing Candidate toward the Small Magellanic Cloud. {\em ApJ, 491}, L11-L13
\par\noindent\hangindent=1pc\hangafter=1 Alcock, C., Allsman, R.~A., Alves, D.~R., Axelrod, T.~S., Becker,
A.~C., Bennett, D.~P., Cook, K.~H., Dalal, N., Drake, A.~J., Freeman,
K.~C., Geha, M., Griest, K., Lehner, M.~J., Marshall, S.~L., Minniti, D.,
Nelson, C.~A., Peterson, B.~A., Popowski, P., Pratt, M.~R., Quinn, P.~J.,
Stubbs, C.~W., Sutherland, W., Tomaney, A.~B., Vandehei, T.,
\& Welch, D.\ (2000) The MACHO Project: Microlensing Results from 5.7 Years of Large Magellanic Cloud Observations. {\em ApJ, 542}, 281-307
\par\noindent\hangindent=1pc\hangafter=1 Agol, E.\ (2002) Occultation and Microlensing. {\em ApJ, 579},
430-436
\par\noindent\hangindent=1pc\hangafter=1 An, J.~H.\ (2005) Gravitational lens under perturbations: symmetry of
perturbing potentials with invariant caustics. {\em MNRAS, 356},
1409-1428
\par\noindent\hangindent=1pc\hangafter=1 An, J.~H.,
\& Evans, N.~W.\ (2006) The Chang-Refsdal lens revisited. {\em MNRAS, 369}, 317-334
\par\noindent\hangindent=1pc\hangafter=1 Aubourg, E., Bareyre, P., Brehin, S., Gros, M., Lachieze-Rey, M.,
Laurent, B., Lesquoy, E., Magneville, C., Milsztajn, A., Moscoso, L.,
Queinnec, F., Rich, J., Spiro, M., Vigroux, L., Zylberajch, S., Ansari, R.,
Cavalier, F., Moniez, M., Beaulieu, J.~P., Ferlet, R., Grison, P.,
Vidal-Madjar, A., Guibert, J., Moreau, O., Tajahmady, F., Maurice, E.,
Prevot, L.,
\& Gry, C.\ (1993) Evidence for Gravitational Microlensing by Dark Objects in the Galactic Halo. {\em Nature, 365}, 623-625
\par\noindent\hangindent=1pc\hangafter=1 Beaulieu, J.-P., Bennett, D.~P., Fouqu{\'e}, P., Williams, A.,
Dominik, M., J{\o}rgensen, U.~G., Kubas, D., Cassan, A., Coutures, C.,
Greenhill, J., Hill, K., Menzies, J., Sackett, P.~D., Albrow, M., Brillant,
S., Caldwell, J.~A.~R., Calitz, J.~J., Cook, K.~H., Corrales, E., Desort,
M., Dieters, S., Dominis, D., Donatowicz, J., Hoffman, M., Kane, S.,
Marquette, J.-B., Martin, R., Meintjes, P., Pollard, K., Sahu, K., Vinter,
C., Wambsganss, J., Woller, K., Horne, K., Steele, I., Bramich, D.~M.,
Burgdorf, M., Snodgrass, C., Bode, M., Udalski, A., Szyma{\'n}ski, M.~K.,
Kubiak, M., Wi{\c e}ckowski, T., Pietrzy{\'n}ski, G., Soszy{\'n}ski, I.,
Szewczyk, O., Wyrzykowski, {\L}., Paczy{\'n}ski, B., Abe, F., Bond, I.~A.,
Britton, T.~R., Gilmore, A.~C., Hearnshaw, J.~B., Itow, Y., Kamiya, K.,
Kilmartin, P.~M., Korpela, A.~V., Masuda, K., Matsubara, Y., Motomura, M.,
Muraki, Y., Nakamura, S., Okada, C., Ohnishi, K., Rattenbury, N.~J., Sako,
T., Sato, S., Sasaki, M., Sekiguchi, T., Sullivan, D.~J., Tristram, P.~J.,
Yock, P.~C.~M.,
\& Yoshioka, T.\ (2006) Discovery of a cool planet of 5.5 Earth masses through gravitational microlensing. {\em Nature, 439}, 437-440
\par\noindent\hangindent=1pc\hangafter=1 Boutreux, T.,
\& Gould, A.\ (1996) Monte Carlo Simulations of MACHO Parallaxes from a Satellite. {\em ApJ, 462}, 705-
\par\noindent\hangindent=1pc\hangafter=1 Bennett, D.~P.\ (2004) The Detection of Terrestrial Planets via
Gravitational Microlensing: Space vs. Ground-based Surveys. {\em ASPC,
321}, 59-67
\par\noindent\hangindent=1pc\hangafter=1 Bennett, D.~P.\ (2009a) Detection of Extrasolar Planets by
Gravitational Microlensing. {\it Exoplanets: Detection, Formation, Properties, Habitability},
eds.\ J.\ Mason, (Berlin:Springer) (arXiv:0902.1761)
\par\noindent\hangindent=1pc\hangafter=1 Bennett, D.~P.\ (2009b) An Efficient Method for Modeling High
Magnification Planetary Microlensing Events. {\em ApJ}, submitted (arXiv:0911.2703)
\par\noindent\hangindent=1pc\hangafter=1 Bennett, D.~P., Anderson, J.,
\& Gaudi, B.~S.\ (2007) Characterization of Gravitational Microlensing Planetary Host Stars. {\em ApJ, 660}, 781-790
\par\noindent\hangindent=1pc\hangafter=1 Bennett, D.~P.,
\& Rhie, S.~H.\ (1996) Detecting Earth-Mass Planets with Gravitational Microlensing. {\em ApJ, 472}, 660-664
\par\noindent\hangindent=1pc\hangafter=1 Bennett, D.~P.,
\& Rhie, S.~H.\ (2002) Simulation of a Space-based Microlensing Survey for Terrestrial Extrasolar Planets. {\em ApJ, 574}, 985-1003
\par\noindent\hangindent=1pc\hangafter=1 Bennett, D.~P., Anderson, J., Bond, I.~A., Udalski, A.,
\& Gould, A.\ (2006) Identification of the OGLE-2003-BLG-235/MOA-2003-BLG-53 Planetary Host Star. {\em ApJ, 647}, L171-L174
\par\noindent\hangindent=1pc\hangafter=1 Bennett, D.~P., Anderson, J.,
\& Gaudi, B.~S.\ (2007a) Characterization of Gravitational Microlensing Planetary Host Stars. {\em ApJ, 660}, 781-790
\par\noindent\hangindent=1pc\hangafter=1 Bennett, D.~P., Anderson, J., Beaulieu, J.~P., Bond, I., Cheng, E.,
Cook, K., Friedman, S., Gaudi, B.~S., Gould, A., Jenkins, J., Kimble, R.,
Lin, D., Mather, J., Rich, M., Sahu, K., Sumi, T., Tenerelli, D., Udalski,
A.,
\& Yoch, P.\ (2009) A Census of Explanets in Orbits Beyond 0.5 AU via Space-based Microlensing. {\em Astro2010: The Astronomy and Astrophysics Decadal Survey, Science White Papers}, 18
\par\noindent\hangindent=1pc\hangafter=1 Bennett, D.~P., Bond, I.~A., Udalski, A., Sumi, T., Abe, F., Fukui,
A., Furusawa, K., Hearnshaw, J.~B., Holderness, S., Itow, Y., Kamiya, K.,
Korpela, A.~V., Kilmartin, P.~M., Lin, W., Ling, C.~H., Masuda, K.,
Matsubara, Y., Miyake, N., Muraki, Y., Nagaya, M., Okumura, T., Ohnishi,
K., Perrott, Y.~C., Rattenbury, N.~J., Sako, T., Saito, T., Sato, S.,
Skuljan, L., Sullivan, D.~J., Sweatman, W.~L., Tristram, P.~J., Yock,
P.~C.~M., Kubiak, M., Szyma{\'n}ski, M.~K., Pietrzy{\'n}ski, G.,
Soszy{\'n}ski, I., Szewczyk, O., Wyrzykowski, {\L}., Ulaczyk, K., Batista,
V., Beaulieu, J.~P., Brillant, S., Cassan, A., Fouqu{\'e}, P., Kervella,
P., Kubas, D.,
\& Marquette, J.~B.\ (2008) A Low-Mass Planet with a Possible Sub-Stellar-Mass Host in Microlensing Event MOA-2007-BLG-192. {\em ApJ, 684}, 663-683
\par\noindent\hangindent=1pc\hangafter=1 Bennett, D.~P., Rhie, S.~H., Nikolaev, S., Gaudi, B.~S., Udalski, A.,
Gould, A., Christie, G.~W., Maoz, D., Dong, S., McCormick, J., Szymanski,
M.~K., Tristram, P.~J., Macintosh, B., Cook, K.~H., Kubiak, M.,
Pietrzynski, G., Soszynski, I., Szewczyk, O., Ulaczyk, K., Wyrzykowski, L.,
DePoy, D.~L., Han, C., Kaspi, S., Lee, C.~-., Mallia, F., Natusch, T.,
Park, B.~-., Pogge, R.~W., Polishook, D., Abe, F., Bond, I.~A., Botzler,
C.~S., Fukui, A., Hearnshaw, J.~B., Itow, Y., Kamiya, K., Korpela, A.~V.,
Kilmartin, P.~M., Lin, W., Masuda, K., Matsubara, Y., Motomura, M., Muraki,
Y., Nakamura, S., Okumura, T., Ohnishi, K., Perrott, Y.~C., Rattenbury,
N.~J., Sako, T., Saito, T., Sato, S., Skuljan, L., Sullivan, D.~J., Sumi,
T., Sweatman, W.~L., Yock, P.~C.~M., Albrow, M., Allan, A., Beaulieu,
J.~-., Bramich, D.~M., Burgdorf, M.~J., Coutures, C., Dominik, M., Dieters,
S., Fouque, P., Greenhill, J., Horne, K., Snodgrass, C., Steele, I.,
Tsapras, Y., Chaboyer, B., Crocker, A.,
\& Frank, S.\ (2009) Masses and Orbital Constraints for the OGLE-2006-BLG-109Lb,c Jupiter/Saturn Analog Planetary System. {\em ApJ}, submitted
(arXiv:0911.2706)
\par\noindent\hangindent=1pc\hangafter=1 Bolatto, A.~D.,
\& Falco, E.~E.\ (1994) The detectability of planetary companions of compact Galactic objects from their effects on microlensed light curves of distant stars. {\em ApJ, 436}, 112-116
\par\noindent\hangindent=1pc\hangafter=1 Bond, I.~A., Udalski, A., Jaroszy{\'n}ski, M., Rattenbury, N.~J.,
Paczy{\'n}ski, B., Soszy{\'n}ski, I., Wyrzykowski, L., Szyma{\'n}ski,
M.~K., Kubiak, M., Szewczyk, O., {\.Z}ebru{\'n}, K., Pietrzy{\'n}ski, G.,
Abe, F., Bennett, D.~P., Eguchi, S., Furuta, Y., Hearnshaw, J.~B., Kamiya,
K., Kilmartin, P.~M., Kurata, Y., Masuda, K., Matsubara, Y., Muraki, Y.,
Noda, S., Okajima, K., Sako, T., Sekiguchi, T., Sullivan, D.~J., Sumi, T.,
Tristram, P.~J., Yanagisawa, T.,
\& Yock, P.~C.~M.\ (2004) OGLE 2003-BLG-235/MOA 2003-BLG-53: A Planetary Microlensing Event. {\em ApJ, 606}, L155-L158
\par\noindent\hangindent=1pc\hangafter=1 Borucki, W.~J., Koch, D.~G., Lissauer, J.~J., Basri, G.~B., Caldwell,
J.~F., Cochran, W.~D., Dunham, E.~W., Geary, J.~C., Latham, D.~W.,
Gilliland, R.~L., Caldwell, D.~A., Jenkins, J.~M.,
\& Kondo, Y.\ (2003) The Kepler mission: a wide-field-of-view photometer designed to determine the frequency of Earth-size planets around solar-like stars. {\em SPIE, 4854}, 129-140
\par\noindent\hangindent=1pc\hangafter=1 Bozza, V.\ (2000) Caustics in special multiple lenses. {\em
A\&A, 355}, 423-432
\par\noindent\hangindent=1pc\hangafter=1 Bromley, B.~C.\ (1996) Finite-Size Gravitational Microlenses. {\em
ApJ, 467}, 537-539
\par\noindent\hangindent=1pc\hangafter=1 Calchi Novati, S., Paulin-Henriksson, S., An, J., Baillon, P.,
Belokurov, V., Carr, B.~J., Cr{\'e}z{\'e}, M., Evans, N.~W.,
Giraud-H{\'e}raud, Y., Gould, A., Hewett, P., Jetzer, P., Kaplan, J.,
Kerins, E., Smartt, S.~J., Stalin, C.~S., Tsapras, Y.,
\& Weston, M.~J.\ (2005) POINT-AGAPE pixel lensing survey of M 31. Evidence for a MACHO contribution to galactic halos. {\em A\&A, 443}, 911-928
\par\noindent\hangindent=1pc\hangafter=1 Cassan, A.\ (2008) An alternative parameterisation for binary-lens
caustic-crossing events. {\em A\&A, 491}, 587-595
\par\noindent\hangindent=1pc\hangafter=1 Chang, K.,
\& Refsdal, S.\ (1979) Flux variations of QSO 0957+561 A, B and image splitting by stars near the light path. {\em Nature, 282}, 561-564
\par\noindent\hangindent=1pc\hangafter=1 Chang, K.,
\& Refsdal, S.\ (1984) Star disturbances in gravitational lens galaxies. {\em A\&A, 132}, 168-178
\par\noindent\hangindent=1pc\hangafter=1 Chung, S.-J., Han, C., Park, B.-G., Kim, D., Kang, S., Ryu, Y.-H.,
Kim, K.~M., Jeon, Y.-B., Lee, D.-W., Chang, K., Lee, W.-B.,
\& Kang, Y.~H.\ (2005) Properties of Central Caustics in Planetary Microlensing. {\em ApJ, 630}, 535-542
\par\noindent\hangindent=1pc\hangafter=1 Chung, S.-J., Kim, D., Darnley, M.~J., Duke, J.~P., Gould, A., Han,
C., Jeon, Y.-B., Kerins, E., Newsam, A.,
\& Park, B.-G.\ (2006) The Possibility of Detecting Planets in the Andromeda Galaxy. {\em ApJ, 650}, 432-437
\par\noindent\hangindent=1pc\hangafter=1 Covone, G., de Ritis, R., Dominik, M.,
\& Marino, A.~A.\ (2000) Detecting planets around stars in nearby galaxies. {\em A\&A, 357}, 816-822
\par\noindent\hangindent=1pc\hangafter=1 de Jong, J.~T.~A., Kuijken, K., Crotts, A.~P.~S., Sackett, P.~D.,
Sutherland, W.~J., Uglesich, R.~R., Baltz, E.~A., Cseresnjes, P., Gyuk, G.,
Widrow, L.~M.,
\& The MEGA collaboration (2004) First microlensing candidates from the MEGA survey of M 31. {\em A\&A, 417}, 461-477
\par\noindent\hangindent=1pc\hangafter=1 Derue, F., Afonso, C., Alard, C., Albert, J.-N., Andersen, J.,
Ansari, R., Aubourg, {\'E}., Bareyre, P., Bauer, F., Beaulieu, J.-P.,
Blanc, G., Bouquet, A., Char, S., Charlot, X., Couchot, F., Coutures, C.,
Ferlet, R., Fouqu{\'e}, P., Glicenstein, J.-F., Goldman, B., Gould, A.,
Graff, D., Gros, M., Ha{\"i}ssinski, J., Hamilton, J.-C., Hardin, D., de
Kat, J., Kim, A., Lasserre, T., Le Guillou, L., Lesquoy, {\'E}., Loup, C.,
Magneville, C., Mansoux, B., Marquette, J.-B., Maurice, {\'E}., Milsztajn,
A., Moniez, M., Palanque-Delabrouille, N., Perdereau, O., Pr{\'e}vot, L.,
Regnault, N., Rich, J., Spiro, M., Vidal-Madjar, A., Vigroux, L.,
\& Zylberajch, S.\ (2001) Observation of microlensing toward the galactic spiral arms. EROS II 3 year survey. {\em A\&A, 373}, 126-138
\par\noindent\hangindent=1pc\hangafter=1 Di Stefano, R.,
\& Perna, R.\ (1997) Identifying Microlensing by Binaries. {\em ApJ, 488}, 55-63
\par\noindent\hangindent=1pc\hangafter=1 Di Stefano, R.\ (1999) Microlensing and the Search for
Extraterrestrial Life. {\em ApJ, 512}, 558-563
\par\noindent\hangindent=1pc\hangafter=1 Di Stefano, R.,
\& Scalzo, R.~A.\ (1999a) A New Channel for the Detection of Planetary Systems through Microlensing. I. Isolated Events due to Planet Lenses. {\em ApJ, 512}, 564-578
\par\noindent\hangindent=1pc\hangafter=1 Di Stefano, R.,
\& Scalzo, R.~A.\ (1999b) A New Channel for the Detection of Planetary Systems through Microlensing. II. Repeating Events. {\em ApJ, 512}, 579-600
\par\noindent\hangindent=1pc\hangafter=1 Di Stefano, R.\ (2008) Mesolensing Explorations of Nearby Masses:
From Planets to Black Holes. {\em ApJ, 684}, 59-67
\par\noindent\hangindent=1pc\hangafter=1 Di Stefano, R.,
\& Night, C.\ (2008) Discovery and Study of Nearby Habitable Planets with Mesolensing. {\em ApJ}, submitted (arXiv:0801.1510)
\par\noindent\hangindent=1pc\hangafter=1 Dominik, M.\ (1995) Improved Routines for the Inversion of the
Gravitational Lens Equation for a Set of Source Points. {\em
A\&AS, 109}, 597-610
\par\noindent\hangindent=1pc\hangafter=1 Dominik, M.,
\& Hirshfeld, A.~C.\ (1996) Evidence for a binary lens in the MACHO LMC No. 1 microlensing event. {\em A\&A, 313}, 841-850
\par\noindent\hangindent=1pc\hangafter=1 Dominik, M.\ (1998) Galactic microlensing with rotating binaries.
{\em A\&A, 329}, 361-374
\par\noindent\hangindent=1pc\hangafter=1 Dominik, M.\ (1999a) Ambiguities in FITS of observed binary lens
galactic microlensing events. {\em A\&A, 341}, 943-953
\par\noindent\hangindent=1pc\hangafter=1 Dominik, M.\ (1999b) The binary gravitational lens and its extreme
cases. {\em A\&A, 349}, 108-125
\par\noindent\hangindent=1pc\hangafter=1 Dominik, M.\ (2004a) Theory and practice of microlensing light curves
around fold singularities. {\em MNRAS, 353}, 69-86
\par\noindent\hangindent=1pc\hangafter=1 Dominik, M.\ (2004b) Revealing stellar brightness profiles by means of
microlensing fold caustics. {\em MNRAS, 353}, 118-132
\par\noindent\hangindent=1pc\hangafter=1 Dominik, M.\ (2006) Stochastic distributions of lens and source
properties for observed galactic microlensing events. {\em MNRAS, 367},
669-692
\par\noindent\hangindent=1pc\hangafter=1 Dominik, M.\ (2007) Adaptive contouring - an efficient way to
calculate microlensing light curves of extended sources. {\em MNRAS,
377}, 1679-1688
\par\noindent\hangindent=1pc\hangafter=1 Dominik, M., Jorgensen, U.~G., Horne, K., Tsapras, Y., Street, R.~A.,
Wyrzykowski, L., Hessman, F.~V., Hundertmark, M., Rahvar, S., Wambsganss,
J., Scarpetta, G., Bozza, V., Calchi Novati, S., Mancini, L., Masi, G.,
Teuber, J., Hinse, T.~C., Steele, I.~A., Burgdorf, M.~J.,
\& Kane, S.\ (2008) Inferring statistics of planet populations by means of automated microlensing searches.
(arXiv:0808.0004)
\par\noindent\hangindent=1pc\hangafter=1 Dominik, M.\ (2009) Parameter degeneracies and (un)predictability of
gravitational microlensing events. {\em MNRAS, 393}, 816-821
\par\noindent\hangindent=1pc\hangafter=1 Dong, S., DePoy, D.~L., Gaudi, B.~S., Gould, A., Han, C., Park,
B.-G., Pogge, R.~W., Udalski, A., Szewczyk, O., Kubiak, M., Szyma{\'n}ski,
M.~K., Pietrzy{\'n}ski, G., Soszy{\'n}ski, I., Wyrzykowski, {\L}., Zebru{\'n}, K.\ (2006) Planetary Detection Efficiency of the Magnification 3000 Microlensing Event OGLE-2004-BLG-343. {\em ApJ, 642}, 842-860
\par\noindent\hangindent=1pc\hangafter=1 Dong, S., Udalski, A., Gould, A., Reach, W.~T., Christie, G.~W.,
Boden, A.~F., Bennett, D.~P., Fazio, G., Griest, K., Szyma{\'n}ski, M.~K.,
Kubiak, M., Soszy{\'n}ski, I., Pietrzy{\'n}ski, G., Szewczyk, O.,
Wyrzykowski, {\L}., Ulaczyk, K., Wieckowski, T., Paczy{\'n}ski, B., DePoy,
D.~L., Pogge, R.~W., Preston, G.~W., Thompson, I.~B.,
\& Patten, B.~M.\ (2007) First Space-Based Microlens Parallax Measurement: Spitzer Observations of OGLE-2005-SMC-001. {\em ApJ, 664}, 862-878
\par\noindent\hangindent=1pc\hangafter=1 Dong, S., Gould, A., Udalski, A., Anderson, J., Christie, G.~W.,
Gaudi, B.~S., The OGLE Collaboration, Jaroszy{\'n}ski, M., Kubiak, M.,
Szyma{\'n}ski, M.~K., Pietrzy{\'n}ski, G., Soszy{\'n}ski, I., Szewczyk, O.,
Ulaczyk, K., Wyrzykowski, {\L}., The {$\mu$}FUN Collaboration, DePoy,
D.~L., Fox, D.~B., Gal-Yam, A., Han, C., L{\'e}pine, S., McCormick, J.,
Ofek, E., Park, B.-G., Pogge, R.~W., The MOA Collaboration, Abe, F.,
Bennett, D.~P., Bond, I.~A., Britton, T.~R., Gilmore, A.~C., Hearnshaw,
J.~B., Itow, Y., Kamiya, K., Kilmartin, P.~M., Korpela, A., Masuda, K.,
Matsubara, Y., Motomura, M., Muraki, Y., Nakamura, S., Ohnishi, K., Okada,
C., Rattenbury, N., Saito, T., Sako, T., Sasaki, M., Sullivan, D., Sumi,
T., Tristram, P.~J., Yanagisawa, T., Yock, P.~C.~M., Yoshoika, T., The
PLANET/Robo Net Collaborations, Albrow, M.~D., Beaulieu, J.~P., Brillant,
S., Calitz, H., Cassan, A., Cook, K.~H., Coutures, C., Dieters, S.,
Prester, D.~D., Donatowicz, J., Fouqu{\'e}, P., Greenhill, J., Hill, K.,
Hoffman, M., Horne, K., J{\o}rgensen, U.~G., Kane, S., Kubas, D.,
Marquette, J.~B., Martin, R., Meintjes, P., Menzies, J., Pollard, K.~R.,
Sahu, K.~C., Vinter, C., Wambsganss, J., Williams, A., Bode, M., Bramich,
D.~M., Burgdorf, M., Snodgrass, C., Steele, I., Doublier, V.,
\& Foellmi, C.\ (2009a) OGLE-2005-BLG-071Lb, the Most Massive M Dwarf Planetary Companion? {\em ApJ, 695}, 970-987
\par\noindent\hangindent=1pc\hangafter=1 Dong, S., Bond, I.~A., Gould, A., Koz{\l}owski, S., Miyake, N.,
Gaudi, B.~S., Bennett, D.~P., Abe, F., Gilmore, A.~C., Fukui, A., Furusawa,
K., Hearnshaw, J.~B., Itow, Y., Kamiya, K., Kilmartin, P.~M., Korpela, A.,
Lin, W., Ling, C.~H., Masuda, K., Matsubara, Y., Muraki, Y., Nagaya, M.,
Ohnishi, K., Okumura, T., Perrott, Y.~C., Rattenbury, N., Saito, T., Sako,
T., Sato, S., Skuljan, L., Sullivan, D.~J., Sumi, T., Sweatman, W.,
Tristram, P.~J., Yock, P.~C.~M., The MOA Collaboration, Bolt, G., Christie,
G.~W., DePoy, D.~L., Han, C., Janczak, J., Lee, C.-U., Mallia, F.,
McCormick, J., Monard, B., Maury, A., Natusch, T., Park, B.-G., Pogge,
R.~W., Santallo, R., Stanek, K.~Z., The {$\mu$}FUN Collaboration, Udalski,
A., Kubiak, M., Szyma{\'n}ski, M.~K., Pietrzy{\'n}ski, G., Soszy{\'n}ski,
I., Szewczyk, O., Wyrzykowski, {\L}., Ulaczyk, K.,
\& The OGLE Collaboration (2009b) Microlensing Event MOA-2007-BLG-400: Exhuming the Buried Signature of a Cool, Jovian-Mass Planet. {\em ApJ, 698}, 1826-1837
\par\noindent\hangindent=1pc\hangafter=1 Einstein, A.\ (1936) Lens-Like Action of a Star by the Deviation of
Light in the Gravitational Field. {\em Sci, 84}, 506-507
\par\noindent\hangindent=1pc\hangafter=1 Fields, D.~L., Albrow, M.~D., An, J., Beaulieu, J.-P., Caldwell,
J.~A.~R., DePoy, D.~L., Dominik, M., Gaudi, B.~S., Gould, A., Greenhill,
J., Hill, K., J{\o}rgensen, U.~G., Kane, S., Martin, R., Menzies, J.,
Pogge, R.~W., Pollard, K.~R., Sackett, P.~D., Sahu, K.~C., Vermaak, P.,
Watson, R., Williams, A., Glicenstein, J.-F.,
\& Hauschildt, P.~H.\ (2003) High-Precision Limb-Darkening Measurement of a K3 Giant Using Microlensing. {\em ApJ, 596}, 1305-1319
\par\noindent\hangindent=1pc\hangafter=1 Fluke, C.~J.,
\& Webster, R.~L.\ (1999) Investigating the geometry of quasars with microlensing. {\em MNRAS, 302}, 68-74
\par\noindent\hangindent=1pc\hangafter=1 Ford, E.~B.,
\& Rasio, F.~A.\ (2008) Origins of Eccentric Extrasolar Planets: Testing the Planet-Planet Scattering Model. {\em ApJ, 686}, 621-636
\par\noindent\hangindent=1pc\hangafter=1 Gaudi, B.~S.\ (1998) Distinguishing Between Binary-Source and
Planetary Microlensing Perturbations. {\em ApJ, 506}, 533-539
\par\noindent\hangindent=1pc\hangafter=1 Gaudi, B.~S.,
\& Gould, A.\ (1997a) Satellite Parallaxes of Lensing Events toward the Galactic Bulge. {\em ApJ, 477}, 152-162
\par\noindent\hangindent=1pc\hangafter=1 Gaudi, B.~S.,
\& Gould, A.\ (1997b) Planet Parameters in Microlensing Events. {\em ApJ, 486}, 85-99
\par\noindent\hangindent=1pc\hangafter=1 Gaudi, B.~S.,
\& Gould, A.\ (1999) Spectrophotometric Resolution of Stellar Surfaces with Microlensing. {\em ApJ, 513}, 619-625
\par\noindent\hangindent=1pc\hangafter=1 Gaudi, B.~S.,
\& Han, C.\ (2004) The Many Possible Interpretations of Microlensing Event OGLE 2002-BLG-055. {\em ApJ, 611}, 528-536
\par\noindent\hangindent=1pc\hangafter=1 Gaudi et al.(1998) Microlensing by Multiple Planets in High-Magnification Events. {\em ApJ, 502}, L33-L37
\par\noindent\hangindent=1pc\hangafter=1 Gaudi, B.~S.,
\& Petters, A.~O.\ (2002) Gravitational Microlensing near Caustics. I. Folds. {\em ApJ, 574}, 970-984
\par\noindent\hangindent=1pc\hangafter=1 Gaudi, B.~S.,
\& Petters, A.~O.\ (2002) Gravitational Microlensing near Caustics. II. Cusps. {\em ApJ, 580}, 468-489
\par\noindent\hangindent=1pc\hangafter=1 Gaudi, B.~S., Albrow, M.~D., An, J., Beaulieu, J.-P., Caldwell,
J.~A.~R., DePoy, D.~L., Dominik, M., Gould, A., Greenhill, J., Hill, K.,
Kane, S., Martin, R., Menzies, J., Naber, R.~M., Pel, J.-W., Pogge, R.~W.,
Pollard, K.~R., Sackett, P.~D., Sahu, K.~C., Vermaak, P., Vreeswijk, P.~M.,
Watson, R.,
\& Williams, A.\ (2002) Microlensing Constraints on the Frequency of Jupiter-Mass Companions: Analysis of 5 Years of PLANET Photometry. {\em ApJ, 566}, 463-499
\par\noindent\hangindent=1pc\hangafter=1 Gaudi, B.~S., Bennett, D.~P., Udalski, A., Gould, A., Christie,
G.~W., Maoz, D., Dong, S., McCormick, J., Szyma{\'n}ski, M.~K., Tristram,
P.~J., Nikolaev, S., Paczy{\'n}ski, B., Kubiak, M., Pietrzy{\'n}ski, G.,
Soszy{\'n}ski, I., Szewczyk, O., Ulaczyk, K., Wyrzykowski, {\L}., DePoy, ould
D.~L., Han, C., Kaspi, S., Lee, C.-U., Mallia, F., Natusch, T., Pogge,
R.~W., Park, B.-G., Abe, F., Bond, I.~A., Botzler, C.~S., Fukui, A.,
Hearnshaw, J.~B., Itow, Y., Kamiya, K., Korpela, A.~V., Kilmartin, P.~M.,
Lin, W., Masuda, K., Matsubara, Y., Motomura, M., Muraki, Y., Nakamura, S.,
Okumura, T., Ohnishi, K., Rattenbury, N.~J., Sako, T., Saito, T., Sato, S.,
Skuljan, L., Sullivan, D.~J., Sumi, T., Sweatman, W.~L., Yock, P.~C.~M.,
Albrow, M.~D., Allan, A., Beaulieu, J.-P., Burgdorf, M.~J., Cook, K.~H.,
Coutures, C., Dominik, M., Dieters, S., Fouqu{\'e}, P., Greenhill, J.,
Horne, K., Steele, I., Tsapras, Y., Chaboyer, B., Crocker, A., Frank, S.,
\& Macintosh, B.\ (2008a) Discovery of a Jupiter/Saturn Analog with Gravitational Microlensing. {\em Sci, 319}, 927-930
\par\noindent\hangindent=1pc\hangafter=1 Gaudi, B.~S., Patterson, J., Spiegel, D.~S., Krajci, T., Koff, R.,
Pojma{\'n}ski, G., Dong, S., Gould, A., Prieto, J.~L., Blake, C.~H.,
Roming, P.~W.~A., Bennett, D.~P., Bloom, J.~S., Boyd, D., Eyler, M.~E., de
Ponthi{\`e}re, P., Mirabal, N., Morgan, C.~W., Remillard, R.~R.,
Vanmunster, T., Wagner, R.~M.,
\& Watson, L.~C.\ (2008b) Discovery of a Very Bright, Nearby Gravitational Microlensing Event. {\em ApJ, 677}, 1268-1277
\par\noindent\hangindent=1pc\hangafter=1 Goldreich, P., Lithwick, Y.,
\& Sari, R.\ (2004) Final Stages of Planet Formation. {\em ApJ, 614}, 497-507
\par\noindent\hangindent=1pc\hangafter=1 Gould, A.\ (1992) Extending the MACHO search to about 10$^6$ solar
masses. {\em ApJ, 392}, 442-451
\par\noindent\hangindent=1pc\hangafter=1 Gould, A.\ (1994) Proper motions of MACHOs. {\em ApJ, 421}, L71-L74
\par\noindent\hangindent=1pc\hangafter=1 Gould, A.\ (1994) MACHO velocities from satellite-based parallaxes.
{\em ApJ, 421}, L75-L78
\par\noindent\hangindent=1pc\hangafter=1 Gould, A.\ (1995) MACHO parallaxes from a single satellite. {\em
ApJ, 441}, L21-L24
\par\noindent\hangindent=1pc\hangafter=1 Gould, A.\ (1996) Microlensing and the Stellar Mass Function. {\em
PASP, 108}, 465-476
\par\noindent\hangindent=1pc\hangafter=1 Gould, A.\ (1999) Microlens Parallaxes with SIRTF. {\em ApJ, 514},
869-877
\par\noindent\hangindent=1pc\hangafter=1 Gould, A.\ (2000) Measuring the Remnant Mass Function of the Galactic
Bulge. {\em ApJ, 535}, 928-931
\par\noindent\hangindent=1pc\hangafter=1 Gould, A.\ (2004) Resolution of the MACHO-LMC-5 Puzzle: The
Jerk-Parallax Microlens Degeneracy. {\em ApJ, 606}, 319-325
\par\noindent\hangindent=1pc\hangafter=1 Gould, A.\ (2008) Hexadecapole Approximation in Planetary
Microlensing. {\em ApJ, 681}, 1593-1598
\par\noindent\hangindent=1pc\hangafter=1 Gould, A.\ (2009) Wide Field Imager in Space for Dark Energy and
Planets. {\em Astro2010: The Astronomy and Astrophysics Decadal Survey, Science White Papers}, 100
(arXiv:0902.2211)
\par\noindent\hangindent=1pc\hangafter=1 Gould, A.,
\& Gaucherel, C.\ (1997) Stokes's Theorem Applied to Microlensing of Finite Sources. {\em ApJ, 477}, 580-584
\par\noindent\hangindent=1pc\hangafter=1 Gould, A.,
\& Loeb, A.\ (1992) Discovering planetary systems through gravitational microlenses. {\em ApJ, 396}, 104-114
\par\noindent\hangindent=1pc\hangafter=1 Gould, A.,
\& Han, C.\ (2000) Astrometric Resolution of Severely Degenerate Binary Microlensing Events. {\em ApJ, 538}, 653-656
\par\noindent\hangindent=1pc\hangafter=1 Gould, A., Miralda-Escude, J.,
\& Bahcall, J.~N.\ (1994) Microlensing Events: Thin Disk, Thick Disk, or Halo? {\em ApJ, 423}, L105-108
\par\noindent\hangindent=1pc\hangafter=1 Gould, A., Udalski, A., An, D., Bennett, D.~P., Zhou, A.-Y.,
Dong, S., Rattenbury, N.~J., Gaudi, B.~S., Yock, P.~C.~M., Bond,
I.~A., Christie, G.~W., Horne, K., Anderson, J., Stanek, K.~Z., DePoy,
D.~L., Han, C., McCormick, J., Park, B.-G., Pogge, R.~W., Poindexter,
S.~D., Soszy{\'n}ski, I., Szyma{\'n}ski, M.~K., Kubiak, M.,
Pietrzy{\'n}ski, G., Szewczyk, O., Wyrzykowski, {\L}., Ulaczyk, K.,
Paczy{\'n}ski, B., Bramich, D.~M., Snodgrass, C., Steele, I.~A.,
Burgdorf, M.~J., Bode, M.~F., Botzler, C.~S., Mao, S., \& Swaving,
S.~C.\ (2006) Microlens OGLE-2005-BLG-169 Implies That Cool
Neptune-like Planets Are Common. {\em ApJ, 644}, L37-L40
\par\noindent\hangindent=1pc\hangafter=1 Griest, K.,
\& Hu, W.\ (1992) Effect of binary sources on the search for massive astrophysical compact halo objects via microlensing. {\em ApJ, 397}, 362-380
\par\noindent\hangindent=1pc\hangafter=1 Griest, K.,
\& Safizadeh, N.\ (1998) The Use of High-Magnification Microlensing Events in Discovering Extrasolar Planets. {\em ApJ, 500}, 37-50
\par\noindent\hangindent=1pc\hangafter=1 Han, C.\ (1999) Analytic relations between the observed gravitational
microlensing parameters with and without the effect of blending. {\em
MNRAS, 309}, 373-378
\par\noindent\hangindent=1pc\hangafter=1 Han, C.\ (2006) Properties of Planetary Caustics in Gravitational
Microlensing. {\em ApJ, 638}, 1080-1085
\par\noindent\hangindent=1pc\hangafter=1 Han, C.\ (2008) Microlensing Detections of Moons of Exoplanets. {\em
ApJ, 684}, 684-690
\par\noindent\hangindent=1pc\hangafter=1 Han, C.\ (2009) Distinguishing Between Planetary and Binary
Interpretations of Microlensing Central Perturbations Under the Severe
Finite-Source Effect. {\em ApJ, 691}, L9-L12
\par\noindent\hangindent=1pc\hangafter=1 Han, C., Gaudi, B.~S., An, J.~H.,
\& Gould, A.\ (2005) Microlensing Detection and Characterization of Wide-Separation Planets. {\em ApJ, 618}, 962-972
\par\noindent\hangindent=1pc\hangafter=1 Han, C.,
\& Gaudi, B.~S.\ (2008) A Characteristic Planetary Feature in Double-Peaked, High-Magnification Microlensing Events. {\em ApJ, 689}, 53-58
\par\noindent\hangindent=1pc\hangafter=1 Han, C.,
\& Gould, A.\ (1995) The Mass Spectrum of MACHOs from Parallax Measurements. {\em ApJ, 447}, 53-
\par\noindent\hangindent=1pc\hangafter=1 Han, C.,
\& Gould, A.\ (2003) Stellar Contribution to the Galactic Bulge Microlensing Optical Depth. {\em ApJ, 592}, 172-175
\par\noindent\hangindent=1pc\hangafter=1 Han, C.,
\& Han, W.\ (2002) On the Feasibility of Detecting Satellites of Extrasolar Planets via Microlensing. {\em ApJ, 580}, 490-493
\par\noindent\hangindent=1pc\hangafter=1 Han, C.,
\& Park, M.-G.\ (2002) A New Channel to Search for Extra-Solar Systems with Multiple Planets via Gravitational Microlensing. {\em JKAS, 35}, 35-40
\par\noindent\hangindent=1pc\hangafter=1 Hamadache, C., Le Guillou, L., Tisserand, P., Afonso, C., Albert,
J.~N., Andersen, J., Ansari, R., Aubourg, {\'E}., Bareyre, P., Beaulieu,
J.~P., Charlot, X., Coutures, C., Ferlet, R., Fouqu{\'e}, P., Glicenstein,
J.~F., Goldman, B., Gould, A., Graff, D., Gros, M., Haissinski, J., de Kat,
J., Lesquoy, {\'E}., Loup, C., Magneville, C., Marquette, J.~B., Maurice,
{\'E}., Maury, A., Milsztajn, A., Moniez, M., Palanque-Delabrouille, N.,
Perdereau, O., Rahal, Y.~R., Rich, J., Spiro, M., Vidal-Madjar, A.,
Vigroux, L.,
\& Zylberajch, S.\ (2006) Galactic Bulge microlensing optical depth from EROS-2. {\em A\&A, 454}, 185-199
\par\noindent\hangindent=1pc\hangafter=1 Hardy, S.~J.,
\& Walker, M.~A.\ (1995) Parallax effects in binary microlensing events. {\em MNRAS, 276}, L79-L82
\par\noindent\hangindent=1pc\hangafter=1 Holtzman, J.~A., Watson, A.~M., Baum, W.~A., Grillmair, C.~J., Groth,
E.~J., Light, R.~M., Lynds, R.,
\& O'Neil, E.~J., Jr.\ (1998) The Luminosity Function and Initial Mass Function in the Galactic Bulge. {\em AJ, 115}, 1946-1957
\par\noindent\hangindent=1pc\hangafter=1 Holz, D.~E.,
\& Wald, R.~M.\ (1996) Photon Statistics Limits for Earth-based Parallax Measurements of MACHO Events. {\em ApJ, 471}, 64-67
\par\noindent\hangindent=1pc\hangafter=1 Horne, K., Snodgrass, C.,
\& Tsapras, Y.\ (2009) A metric and optimization scheme for microlens planet searches. {\em MNRAS, 396}, 2087-2102
\par\noindent\hangindent=1pc\hangafter=1 Ida, S., \& Lin, D.~N.~C.\ (2004) Toward a Deterministic Model of Planetary Formation. II. The Formation and Retention of Gas Giant Planets around Stars with a Range of Metallicities. {\em ApJ, 616}, 567-572
\par\noindent\hangindent=1pc\hangafter=1 Ida, S.,
\& Lin, D.~N.~C.\ (2005) Toward a Deterministic Model of Planetary Formation. III. Mass Distribution of Short-Period Planets around Stars of Various Masses. {\em ApJ, 626}, 1045-1060
\par\noindent\hangindent=1pc\hangafter=1 Ingrosso, G., Calchi Novati, S., De Paolis, F., Jetzer, P., Nucita,
A.~A.,
\& Zakharov, A.~F.\ (2009) Pixel-lensing as a way to detect extrasolar planets in M31. {\em MNRAS}, in press
(arXiv:0906.1050)
\par\noindent\hangindent=1pc\hangafter=1 Janczak, J., et al.\ (2009) Sub-Saturn Planet MOA-2008-BLG-310Lb: Likely To
Be in the Galactic Bulge. {\em ApJ}, submitted (arXiv:0908.0529)
\par\noindent\hangindent=1pc\hangafter=1 Jaroszy{\'n}ski, M.,
\& Mao, S.\ (2001) Predicting the second caustic crossing in binary microlensing events. {\em MNRAS, 325}, 1546-1552
\par\noindent\hangindent=1pc\hangafter=1 Johnson, J.~A., Gaudi, B.~S., Sumi, T., Bond, I.~A.,
\& Gould, A.\ (2008) A High-Resolution Spectrum of the Highly Magnified Bulge G Dwarf MOA-2006-BLG-099S. {\em ApJ, 685}, 508-520
\par\noindent\hangindent=1pc\hangafter=1 Juri{\'c}, M.,
\& Tremaine, S.\ (2008) Dynamical Origin of Extrasolar Planet Eccentricity Distribution. {\em ApJ, 686}, 603-620
\par\noindent\hangindent=1pc\hangafter=1 Kennedy, G.~M., Kenyon, S.~J.,
\& Bromley, B.~C.\ (2007) Planet formation around M-dwarfs: the moving snow line and super-Earths. {\em Ap\&SS, 311}, 9-13
\par\noindent\hangindent=1pc\hangafter=1 Kennedy, G.~M.,
\& Kenyon, S.~J.\ (2008) Planet Formation around Stars of Various Masses: The Snow Line and the Frequency of Giant Planets. {\em ApJ, 673}, 502-512
\par\noindent\hangindent=1pc\hangafter=1 Kervella, P., Th{\'e}venin, F., Di Folco, E.,
\& S{\'e}gransan, D.\ (2004) The angular sizes of dwarf stars and subgiants. Surface brightness relations calibrated by interferometry. {\em A\&A, 426}, 297-307
\par\noindent\hangindent=1pc\hangafter=1 Khavinson, D., \& Neumann, G.\ (2006) On the number of zeros of certain rational harmonic functions.
{\em Proc.\ Amer.\ Math.\ Soc., 134}, 1077-1085
\par\noindent\hangindent=1pc\hangafter=1 Kiraga, M.,
\& Paczynski, B.\ (1994) Gravitational microlensing of the Galactic bulge stars. {\em ApJ, 430}, L101-L104
\par\noindent\hangindent=1pc\hangafter=1 Laughlin, G., Bodenheimer, P.,
\& Adams, F.~C.\ (2004) The Core Accretion Model Predicts Few Jovian-Mass Planets Orbiting Red Dwarfs. {\em ApJ, 612}, L73-L76
\par\noindent\hangindent=1pc\hangafter=1 Lecar, M., Podolak, M., Sasselov, D.,
\& Chiang, E.\ (2006) On the Location of the Snow Line in a Protoplanetary Disk. {\em ApJ, 640}, 1115-1118
\par\noindent\hangindent=1pc\hangafter=1 Liebes, S.\ (1964) Gravitational Lenses. {\em PhRv, 133}, 835-844
\par\noindent\hangindent=1pc\hangafter=1 Liebig, C.,
\& Wambsganss, J.\ (2009) Detectability of extrasolar moons as gravitational microlenses. {\em A\&A}, submitted (arXiv:0912.2076)
\par\noindent\hangindent=1pc\hangafter=1 Lissauer, J.~J.\ (1987) Timescales for planetary accretion and the
structure of the protoplanetary disk. {\em Icar, 69}, 249-265
\par\noindent\hangindent=1pc\hangafter=1 Mao, S.\ (1992) Gravitational microlensing by a single star plus
external shear. {\em ApJ, 389}, 63-67
\par\noindent\hangindent=1pc\hangafter=1 Mao, S.\ (2008) Introduction to Gravitational Microlensing\\
(arXiv:0811.0441)
\par\noindent\hangindent=1pc\hangafter=1 Mao, S.,
\& Di Stefano, R.\ (1995) Interpretation of gravitational microlensing by binary systems. {\em ApJ, 440}, 22-27
\par\noindent\hangindent=1pc\hangafter=1 Mao, S.,
\& Paczynski, B.\ (1991) Gravitational microlensing by double stars and planetary systems. {\em ApJ, 374}, L37-L40
\par\noindent\hangindent=1pc\hangafter=1 Minniti, D., Vandehei, T., Cook, K.~H., Griest, K.,
\& Alcock, C.\ (1998) Detection of Lithium in a Main Sequence Bulge Star Using Keck I as a 15M Diameter Telescope. {\em ApJ, 499}, L175-L178
\par\noindent\hangindent=1pc\hangafter=1 Paczynski, B.\ (1986) Gravitational microlensing by the galactic
halo. {\em ApJ, 304}, 1-5
\par\noindent\hangindent=1pc\hangafter=1 Paczynski, B.\ (1991) Gravitational microlensing of the Galactic
bulge stars. {\em ApJ, 371}, L63-L67
\par\noindent\hangindent=1pc\hangafter=1 Paczynski, B.\ (1996) Gravitational Microlensing in the Local Group.
{\em ARA\&A, 34}, 419-460
\par\noindent\hangindent=1pc\hangafter=1 Paczynski, B.,
\& Stanek, K.~Z.\ (1998) Galactocentric Distance with the Optical Gravitational Lensing Experiment and HIPPARCOS Red Clump Stars. {\em ApJ, 494}, L219-222
\par\noindent\hangindent=1pc\hangafter=1 Palanque-Delabrouille, N., Afonso, C., Albert, J.~N., Andersen, J.,
Ansari, R., Aubourg, E., Bareyre, P., Bauer, F., Beaulieu, J.~P., Bouquet,
A., Char, S., Charlot, X., Couchot, F., Coutures, C., Derue, F., Ferlet,
R., Glicenstein, J.~F., Goldman, B., Gould, A., Graff, D., Gros, M.,
Haissinski, J., Hamilton, J.~C., Hardin, D., de Kat, J., Lesquoy, E., Loup,
C., Magneville, C., Mansoux, B., Marquette, J.~B., Maurice, E., Milsztajn,
A., Moniez, M., Perdereau, O., Prevot, L., Renault, C., Rich, J., Spiro,
M., Vidal-Madjar, A., Vigroux, L., Zylberajch, S.,
\& The EROS Collaboration (1998) Microlensing towards the Small Magellanic Cloud EROS 2 first year survey. {\em A\&A, 332}, 1-9
\par\noindent\hangindent=1pc\hangafter=1 Park, B.-G., Jeon, Y.-B., Lee, C.-U.,
\& Han, C.\ (2006) Microlensing Sensitivity to Earth-Mass Planets in the Habitable Zone. {\em ApJ, 643}, 1233-1238
\par\noindent\hangindent=1pc\hangafter=1 Paulin-Henriksson, S., Baillon, P., Bouquet, A., Carr, B.~J.,
Cr{\'e}z{\'e}, M., Evans, N.~W., Giraud-H{\'e}raud, Y., Gould, A., Hewett,
P., Kaplan, J., Kerins, E., Lastennet, E., Le Du, Y., Melchior, A.-L.,
Smartt, S.~J.,
\& Valls-Gabaud, D.\ (2002) A Candidate M31/M32 Intergalactic Microlensing Event. {\em ApJ, 576}, L121-L124
\par\noindent\hangindent=1pc\hangafter=1 Peale, S.~J.\ (2001) Probability of Detecting a Planetary Companion
during a Microlensing Event. {\em ApJ, 552}, 889-911
\par\noindent\hangindent=1pc\hangafter=1 Pejcha, O.,
\& Heyrovsk{\'y}, D.\ (2009) Extended-Source Effect and Chromaticity in Two-Point-Mass Microlensing. {\em ApJ, 690}, 1772-1796
\par\noindent\hangindent=1pc\hangafter=1 Petters, A.~O., Levine, H.,
\& Wambsganss, J.\ (2001), {\em Singularity theory and gravitational lensing}. Birkh{\" a}user, Boston.
\par\noindent\hangindent=1pc\hangafter=1 Poindexter, S., Afonso, C., Bennett, D.~P., Glicenstein, J.-F.,
Gould, A., Szyma{\'n}ski, M.~K.,
\& Udalski, A.\ (2005) Systematic Analysis of 22 Microlensing Parallax Candidates. {\em ApJ, 633}, 914-930
\par\noindent\hangindent=1pc\hangafter=1 Press, W.~H., Teukolsky, S.~A., Vetterling, W.~T.,
\& Flannery, B.~P.\ (1992), {\em Numerical recipes in FORTRAN. The art of scientific computing}.
University Press, Cambridge.
\par\noindent\hangindent=1pc\hangafter=1 Rattenbury, N.~J., Bond, I.~A., Skuljan, J.,
\& Yock, P.~C.~M.\ (2002) Planetary microlensing at high magnification. {\em MNRAS, 335}, 159-169
\par\noindent\hangindent=1pc\hangafter=1 Refsdal, S.\ (1964) The gravitational lens effect. {\em MNRAS,
128}, 295-306
\par\noindent\hangindent=1pc\hangafter=1 Refsdal, S.\ (1966) On the possibility of determining the distances
and masses of stars from the gravitational lens effect. {\em MNRAS, 134},
315-319
\par\noindent\hangindent=1pc\hangafter=1 Renn, J., Sauer, T.,
\& Stachel, J.\ (1997) The origin of gravitational lensing: a postscript to Einstein's 1936 Science paper. {\em Sci, 275}, 184-186
\par\noindent\hangindent=1pc\hangafter=1 Rhie, S.~H.\ (1997) Infimum Microlensing Amplification of the Maximum
Number of Images of n-Point Lens Systems. {\em ApJ, 484}, 63-69
\par\noindent\hangindent=1pc\hangafter=1 Rhie, S.~H.\ (2003) How Cumbersome is a Tenth Order Polynomial?: The Case of Gravitational Triple Lens Equation.
(arXiv:astro-ph/0202294)
\par\noindent\hangindent=1pc\hangafter=1 Rhie, S.~H.\ (2003) n-point Gravitational Lenses with 5(n-1) Images.
(arXiv:astro-ph/0305166)
\par\noindent\hangindent=1pc\hangafter=1 Rhie, S.~H.,
\& Bennett, D.~P.\ (1999) Line Caustic Microlensing and Limb Darkening. (arXiv:astro-ph/9912050)
\par\noindent\hangindent=1pc\hangafter=1 Rhie, S.~H., Bennett, D.~P., Becker, A.~C., Peterson, B.~A., Fragile,
P.~C., Johnson, B.~R., Quinn, J.~L., Crouch, A., Gray, J., King, L.,
Messenger, B., Thomson, S., Bond, I.~A., Abe, F., Carter, B.~S., Dodd,
R.~J., Hearnshaw, J.~B., Honda, M., Jugaku, J., Kabe, S., Kilmartin, P.~M.,
Koribalski, B.~S., Masuda, K., Matsubara, Y., Muraki, Y., Nakamura, T.,
Nankivell, G.~R., Noda, S., Rattenbury, N.~J., Reid, M., Rumsey, N.~J.,
Saito, T., Sato, H., Sato, S., Sekiguchi, M., Sullivan, D.~J., Sumi, T.,
Watase, Y., Yanagisawa, T., Yock, P.~C.~M.,
\& Yoshizawa, M.\ (2000) On Planetary Companions to the MACHO 98-BLG-35 Microlens Star. {\em ApJ, 533}, 378-391
\par\noindent\hangindent=1pc\hangafter=1 Sackett, P.~D.\ (1999) Searching for Unseen Planets via Occultation
and Microlensing. {\it Planets Outside the Solar System: Theory and Observations}, eds.\ J.-M. Mariotti and D. Alloin,
(Boston:Kluwer), 189-228 (arXiv:astro-ph/9811269)
\par\noindent\hangindent=1pc\hangafter=1 Sako, T., Sekiguchi, T., Sasaki, M., Okajima, K., Abe, F., Bond,
I.~A., Hearnshaw, J.~B., Itow, Y., Kamiya, K., Kilmartin, P.~M., Masuda,
K., Matsubara, Y., Muraki, Y., Rattenbury, N.~J., Sullivan, D.~J., Sumi,
T., Tristram, P., Yanagisawa, T.,
\& Yock, P.~C.~M.\ (2008) MOA-cam3: a wide-field mosaic CCD camera for a gravitational microlensing survey in New Zealand. {\em ExA, 22}, 51-66
\par\noindent\hangindent=1pc\hangafter=1 Sauer, T. (2008) Nova Geminorum 1912 and the origin of the idea of gravitational lensing. {\em Arch. Hist. Exact Sci., 62},
1-22
\par\noindent\hangindent=1pc\hangafter=1 Schneider, P.,
\& Weiss, A.\ (1986) The two-point-mass lens - Detailed investigation of a special asymmetric gravitational lens. {\em A\&A, 164}, 237-259
\par\noindent\hangindent=1pc\hangafter=1 Schneider, P.,
\& Weiss, A.\ (1992) The gravitational lens equation near cusps. {\em A\&A, 260}, 1-2
\par\noindent\hangindent=1pc\hangafter=1 Schneider, P., Ehlers, J., \& Falco, E.~E. (1992),
{\em Gravitational Lenses}. Springer-Verlag, Berlin.
\par\noindent\hangindent=1pc\hangafter=1 Smith, M.~C., Mao, S.,
\& Paczy{\'n}ski, B.\ (2003) Acceleration and parallax effects in gravitational microlensing. {\em MNRAS, 339}, 925-936
\par\noindent\hangindent=1pc\hangafter=1 Snodgrass, C., Horne, K.,
\& Tsapras, Y.\ (2004) The abundance of Galactic planets from OGLE-III 2002 microlensing data. {\em MNRAS, 351}, 967-975
\par\noindent\hangindent=1pc\hangafter=1 Sumi, T., Abe, F., Bond, I.~A., Dodd, R.~J., Hearnshaw, J.~B., Honda,
M., Honma, M., Kan-ya, Y., Kilmartin, P.~M., Masuda, K., Matsubara, Y.,
Muraki, Y., Nakamura, T., Nishi, R., Noda, S., Ohnishi, K., Petterson,
O.~K.~L., Rattenbury, N.~J., Reid, M., Saito, T., Saito, Y., Sato, H.,
Sekiguchi, M., Skuljan, J., Sullivan, D.~J., Takeuti, M., Tristram, P.~J.,
Wilkinson, S., Yanagisawa, T.,
\& Yock, P.~C.~M.\ (2003) Microlensing Optical Depth toward the Galactic Bulge from Microlensing Observations in Astrophysics Group Observations during 2000 with Difference Image Analysis. {\em ApJ, 591}, 204-227
\par\noindent\hangindent=1pc\hangafter=1 Sumi, T., Bennett, D.~P., Bond, I.~A., Udalski, A., Batista, V.,
Dominik, M., Fouqu{\'e}, P., Kubas, D., Gould, A., Macintosh, B., Cook, K.,
Dong, S., Skuljan, L., Cassan, A., The MOA Collaboration: F.~Abe, Botzler,
C.~S., Fukui, A., Furusawa, K., Hearnshaw, J.~B., Itow, Y., Kamiya, K.,
Kilmartin, P.~M., Korpela, A., Lin, W., Ling, C.~H., Masuda, K., Matsubara,
Y., Miyake, N., Muraki, Y., Nagaya, M., Nagayama, T., Ohnishi, K., Okumura,
T., Perrott, Y.~C., Rattenbury, N., Saito, T., Sako, T., Sullivan, D.~J.,
Sweatman, W.~L., P., Yock, P.~C.~M., The PLANET Collaboration:
J.~P.~Beaulieu, Cole, A., Coutures, C., Duran, M.~F., Greenhill, J.,
Jablonski, F., Marboeuf, U., Martioli, E., Pedretti, E., Pejcha, O., Rojo,
P., Albrow, M.~D., Brillant, S., Bode, M., Bramich, D.~M., Burgdorf, M.~J.,
Caldwell, J.~A.~R., Calitz, H., Corrales, E., Dieters, S., Dominis Prester,
D., Donatowicz, J., Hill, K., Hoffman, M., Horne, K., J, U.~G., Kains, N.,
Kane, S., Marquette, J.~B., Martin, R., Meintjes, P., Menzies, J., Pollard,
K.~R., Sahu, K.~C., Snodgrass, C., Steele, I., Street, R., Tsapras, Y.,
Wambsganss, J., Williams, A., Zub, M., The OGLE Collaboration: M.~K.~Szyma,
Kubiak, M., Pietrzy, G., Soszy, I., Szewczyk, O., Ulaczyk, K., The
\$$\backslash$mu\$FUN Collaboration: W.~Allen, Christie, G.~W., DePoy,
D.~L., Gaudi, B.~S., Han, C., Janczak, J., Lee, C.~-., McCormick, J.,
Mallia, F., Monard, B., Natusch, T., Park, B.-G., Pogge, R.~W.,
\& Santallo, R.\ (2009) A Cold Neptune-Mass Planet OGLE-2007-BLG-368Lb: Cold Neptunes Are Common. {\em ApJ}, submitted (arXiv:0912.1171)
\par\noindent\hangindent=1pc\hangafter=1 Thomas, C.~L., Griest, K., Popowski, P., Cook, K.~H., Drake, A.~J.,
Minniti, D., Myer, D.~G., Alcock, C., Allsman, R.~A., Alves, D.~R.,
Axelrod, T.~S., Becker, A.~C., Bennett, D.~P., Freeman, K.~C., Geha, M.,
Lehner, M.~J., Marshall, S.~L., Nelson, C.~A., Peterson, B.~A., Quinn,
P.~J., Stubbs, C.~W., Sutherland, W., Vandehei, T.,
\& Welch, D.~L.\ (2005) Galactic Bulge Microlensing Events from the MACHO Collaboration. {\em ApJ, 631}, 906-934
\par\noindent\hangindent=1pc\hangafter=1 Tsapras, Y., Street, R., Horne, K., Snodgrass, C., Dominik, M.,
Allan, A., Steele, I., Bramich, D.~M., Saunders, E.~S., Rattenbury, N.,
Mottram, C., Fraser, S., Clay, N., Burgdorf, M., Bode, M., Lister, T.~A.,
Hawkins, E., Beaulieu, J.~P., Fouqu{\'e}, P., Albrow, M., Menzies, J.,
Cassan, A.,
\& Dominis-Prester, D.\ (2009) RoboNet-II: Follow-up observations of microlensing events with a robotic network of telescopes. {\em AN, 330}, 4-11
\par\noindent\hangindent=1pc\hangafter=1 Udalski, A., Szymanski, M., Kaluzny, J., Kubiak, M., Krzeminski,
W., Mateo, M., Preston, G.~W., \& Paczynski, B.\ (1993) The optical
gravitational lensing experiment. Discovery of the first candidate
microlensing event in the direction of the Galactic Bulge. {\em AcA,
43}, 289-294
\par\noindent\hangindent=1pc\hangafter=1 Udalski, A., Szymanski, M., Kaluzny, J., Kubiak, M., Mateo, M.,
Krzeminski, W., \& Paczynski, B.\ (1994) The Optical Gravitational
Lensing Experiment. The Early Warning System: Real Time Microlensing.
{\em AcA, 44}, 227-234
\par\noindent\hangindent=1pc\hangafter=1 Udalski, A., Zebrun, K., Szymanski, M., Kubiak, M., Pietrzynski,
G., Soszynski, I., \& Wozniak, P.\ (2000) The Optical Gravitational
Lensing Experiment. Catalog of Microlensing Events in the Galactic
Bulge. {\em AcA, 50}, 1-65
\par\noindent\hangindent=1pc\hangafter=1 Udalski, A.\ (2003) The Optical Gravitational Lensing Experiment.
Real Time Data Analysis Systems in the OGLE-III Survey. {\em AcA, 53},
291-305
\par\noindent\hangindent=1pc\hangafter=1 Udalski, A., Jaroszy{\'n}ski, M., Paczy{\'n}ski, B., Kubiak, M.,
Szyma{\'n}ski, M.~K., Soszy{\'n}ski, I., Pietrzy{\'n}ski, G., Ulaczyk, K.,
Szewczyk, O., Wyrzykowski, {\L}., Christie, G.~W., DePoy, D.~L., Dong, S.,
Gal-Yam, A., Gaudi, B.~S., Gould, A., Han, C., L{\'e}pine, S., McCormick,
J., Park, B.-G., Pogge, R.~W., Bennett, D.~P., Bond, I.~A., Muraki, Y.,
Tristram, P.~J., Yock, P.~C.~M., Beaulieu, J.-P., Bramich, D.~M., Dieters,
S.~W., Greenhill, J., Hill, K., Horne, K.,
\& Kubas, D.\ (2005) A Jovian-Mass Planet in Microlensing Event OGLE-2005-BLG-071. {\em ApJ, 628}, L109-L112
\par\noindent\hangindent=1pc\hangafter=1 Uglesich, R.~R., Crotts, A.~P.~S., Baltz, E.~A., de Jong, J., Boyle,
R.~P.,
\& Corbally, C.~J.\ (2004) Evidence of Halo Microlensing in M31. {\em ApJ, 612}, 877-893
\par\noindent\hangindent=1pc\hangafter=1 van Belle, G.~T.\ (1999) Predicting Stellar Angular Sizes. {\em
PASP, 111}, 1515-1523
\par\noindent\hangindent=1pc\hangafter=1 Vermaak, P.\ (2000) The effects of resolved sources and blending on
the detection of planets via gravitational microlensing. {\em MNRAS,
319}, 1011-1019
\par\noindent\hangindent=1pc\hangafter=1 Vermaak, P.\ (2003) Rapid analysis of binary lens gravitational
microlensing light curves. {\em MNRAS, 344}, 651-656
\par\noindent\hangindent=1pc\hangafter=1 Wambsganss, J.\ (1997) Discovering Galactic planets by gravitational
microlensing: magnification patterns and light curves. {\em MNRAS, 284},
172-188
\par\noindent\hangindent=1pc\hangafter=1 Witt, H.~J.\ (1990) Investigaion of high amplification events in
light curves of gravitationally lensed quasars. {\em A\&A, 236}, 311-322
\par\noindent\hangindent=1pc\hangafter=1 Witt, H.~J.\ (1995) The Effect of the Stellar Size on Microlensing at
the Baade Window. {\em ApJ, 449}, 42-46
\par\noindent\hangindent=1pc\hangafter=1 Witt, H.~J.,
\& Mao, S.\ (1994) Can lensed stars be regarded as pointlike for microlensing by MACHOs?. {\em ApJ, 430}, 505-510
\par\noindent\hangindent=1pc\hangafter=1 Witt, H.~J.,
\& Mao, S.\ (1995) On the Minimum Magnification Between Caustic Crossings for Microlensing by Binary and Multiple Stars. {\em ApJ, 447}, L105-L108
\par\noindent\hangindent=1pc\hangafter=1 Wozniak, P.,
\& Paczynski, B.\ (1997) Microlensing of Blended Stellar Images. {\em ApJ, 487}, 55-60
\par\noindent\hangindent=1pc\hangafter=1 Yoo, J., DePoy, D.~L., Gal-Yam, A., Gaudi, B.~S., Gould, A., Han, C.,
Lipkin, Y., Maoz, D., Ofek, E.~O., Park, B.-G., Pogge, R.~W., Udalski, A.,
Soszy{\'n}ski, I., Wyrzykowski, {\L}., Kubiak, M., Szyma{\'n}ski, M.,
Pietrzy{\'n}ski, G., Szewczyk, O., Zebru{\'n}, K.\ (2004) OGLE-2003-BLG-262: Finite-Source Effects from a Point-Mass Lens. {\em ApJ, 603}, 139-151
\par\noindent\hangindent=1pc\hangafter=1 Zakharov, A.~F.\ (1995) On the magnification of gravitational lens
images near cusps.. {\em A\&A, 293}, 1-4
\par\noindent\hangindent=1pc\hangafter=1 Zakharov, A.~F.\ (1999) On the some properties of gravitational lens
equation near cusps. {\em A\&AT, 18}, 17-25
\end{document}
|
\section{Introduction}
Active galactic nuclei (AGN) are bright emitters both at radio and $\gamma$-ray wavelengths. In the 1990s the
Energetic Gamma-Ray Experiment Telescope (EGRET) on board the {\it Compton Gamma-Ray Observatory}
detected over 65 AGN (mostly blazars) at $\gamma$-ray energies.\cite{hartman99,mattox01} In June 2008,
the {\it Fermi Gamma-ray Space Telescope} was launched. Its primary instrument, the Large Area Telescope (LAT), observes
the whole sky mainly in survey mode at energies 100 MeV-300 GeV.\cite{atwood09} During its first
three months of operation, the LAT detected 205 bright $\gamma$-ray sources,\cite{abdo09a} of which
106 were high confidence associations with AGN.\cite{abdo09b} Most of them host bright parsec-scale radio jets.\cite{kovalev09b}
The $\gamma$-ray emission in AGN likely originates in
the relativistic jet, as suggested by the many correlations found between EGRET and radio/mm observations
of AGN. \cite{valtaoja95}$^-$\cite{kovalev05} Further evidence
has recently been found using the Very Long Baseline Array (VLBA) observations by Lister et al.\cite{lister09b} who showed
that the LAT-detected sources have significantly higher apparent jet speeds. It has also been shown that
the brightness temperature of the VLBI core is higher for LAT-detected sources and the $\gamma$-ray photon
flux correlates with the compact radio flux density.\cite{kovalev09}
The LAT-detected sources also have larger apparent opening angles,\cite{pushkarev09} and are more Doppler-boosted.\cite{savolainen09}
Possible links between the radio polarization of jets and $\gamma$-ray flaring has been previously studied,
e.g., by Jorstad et al.\cite{jorstad01a} who established a connection between superluminal component ejections
and $\gamma$-ray flaring observed by EGRET. Using single-dish radio observations from the University of Michigan
Radio Astronomy Observatory, they found that a local maximum in the polarized flux density was observed simultaneously with the
maximum of the $\gamma$-ray emission. Another study by Lister et al.\cite{lister05} compared the
EGRET detections and 15 GHz VLBI linear polarization properties. They did not find differences between the
core properties of EGRET-detected and non-detected objects, although they found indications that the EGRET-detected
objects have higher integrated linear jet polarization. In this paper we will use the better sampled, more complete LAT data
in the same manner and study how the linear polarization properties of the 15 GHz core differ in the LAT-detected
and non-detected objects.
\section{The Sample and Data}
In our analyses we use the flux-density-limited MOJAVE-1 (Monitoring Of Jets in Active galactic nuclei with VLBA Experiments)
sample of highly beamed AGN.\cite{lister09} The sample
consists of all AGN at declinations $\delta > -20^\circ$ which have had a 15 GHz flux density of at least 1.5 Jy (2 Jy at $\delta < 0^\circ$)
at any epoch between 1994.0 and 2004.0. We have only used sources at galactic latitude $|b| > 10^\circ$ following
the definitions of the LAT 3-month bright AGN list,\cite{abdo09b} resulting in a sample of 123 sources.
We have divided our sample into LAT-detected (30 objects) and non-LAT-detected (93 objects) sources based on the
3-month bright AGN list.\cite{abdo09b}
We have used the VLBA 15 GHz total intensity and linear polarization observations since 2002 and determined the polarized core flux
density $P = (Q^2 + U^2)^{1/2}$, where $Q$ and $U$ are the Stokes parameters, fractional polarization $m_c = P/I$,
where $I$ is the total flux density of the core, and
electric vector position angle $\chi = (1/2) \arctan(U/Q)$ at all the epochs where polarization cross-hands were recorded. The core values were defined using the same method as Ref.~\refcite{lister05} by taking the mean over nine contiguous pixels centered at the fitted core position. If the
polarized flux density was less than five times the rms-value in the $U$ and $Q$ parameters, an upper limit was defined to be five times
the rms-value. In these cases EVPA was not calculated.
\section{Polarization Levels}
In order to study if the polarization properties are connected with $\gamma$-ray activity, we have used
polarization observations from August 2008 until August 2009 to roughly match the LAT observation period. The LAT associations
are based on $\gamma$-ray data integrated between August and October 2008.
Figure \ref{f1} shows the distributions of the maximum polarized flux density of LAT-detected and non-LAT-detected sources since August 2008.
Nine of the the non-LAT-detected sources in the smallest bin are upper limits. The medians of the distributions are 45.0 mJy and 21.6 mJy for the LAT-detected and non-detected sources, respectively. The Gehan's Generalized Wilcoxon test
in the ASURV package,\cite{lavalley92} suitable for censored data, gives a
probability p=0.0003 for the distributions of the LAT-detected and non-detected sources to come from the same population.
The result is not surprising considering that
the polarized flux density often follows the total flux density which is also seen to be higher in the LAT-detected sources when measured quasi-simultaneously with the LAT.\cite{kovalev09}
\begin{figure}[pb]
\centerline{\psfig{file=Hovatta_f1.eps,width=9cm}}
\vspace*{8pt}
\caption{Maximum polarized flux density of LAT-detected (shaded) and non-LAT-detected (unshaded) sources since August 2008.\label{f1}}
\end{figure}
We also wanted to see if this is due to variability in the sources and in Fig.~\ref{f2} we show the distributions of median fractional
polarization of LAT-detected and non-LAT-detected sources from 2002 until August 2008 and from August 2008 until August 2009 (LAT-era). The median values of the distributions are 2.14\% and 1.80\% for the pre-LAT values of LAT-detected and non-detected sources,
respectively and 2.50\% and 1.86\% for the values since August 2008. We used the Kolmogorov-Smirnov (K-S) test to see if the LAT-detected
and non-detected sources come from different populations and the non-parametric Mann-Whitney U-test (M-W U-test) to see if the LAT-detected
sources have typically higher values than non-detected ones. No significant differences are seen when the pre-LAT values are used (K-S test p=0.35, M-W U-test p=0.16), yet when the distributions from the LAT-era are studied, the groups differ significantly (K-S test p=0.013, M-W U-test p=0.043).
\begin{figure}[pb]
\centerline{\psfig{file=Hovatta_f2.eps,width=\textwidth}}
\vspace*{8pt}
\caption{Median fractional polarization of LAT-detected (shaded) and non-LAT-detected (unshaded) sources from 2002 until August 2008 (left panel) and since August 2008 (right panel).\label{f2}}
\end{figure}
We calculated the variability indices of fractional polarization (Eq.~\ref{eq1}) and $\chi$ (Eq.~\ref{eq2}) in the method of Refs.~\refcite{aller03} and \refcite{jorstad07}.
\begin{equation}
V_{m_c} = \frac{(m_\mathrm{max} - \sigma_{m_\mathrm{max}}) - (m_\mathrm{min} + \sigma_{m_\mathrm{min}})}{(m_\mathrm{max} - \sigma_{m_\mathrm{max}}) + (m_\mathrm{min} + \sigma_{m_\mathrm{min}})},
\label{eq1}
\end{equation}
where $m_\mathrm{max}$ and $m_\mathrm{min}$ are the maximum and minimum fractional polarization, respectively, measured over all epochs since 2002, and
$\sigma_{m_\mathrm{max}}$ and $\sigma_{m_\mathrm{min}}$ are the corresponding uncertainties.
\begin{equation}
V_{\chi} = \frac{|\Delta\chi| - \sqrt{(\sigma_{\chi_{1}}^2 + \sigma_{\chi_{2}}^2)}}{90},
\label{eq2}
\end{equation}
where $|\Delta\chi|$ is the observed range of polarization angle and $\sigma_{\chi_{1}}$ and $\sigma_{\chi_{2}}$ are the uncertainties in the
two values of EVPAs that define the range.\cite{jorstad07}
Figure \ref{f3} shows the histograms of the variability indices. It can be seen that none of the lowest variability sources are
detected by the LAT. A K-S test indicates significant differences between the distributions, for both variability indices ($p<0.01$ for $V_{m_c}$ and $p<0.001$ for $V_{\chi}$).
\begin{figure}[pb]
\centerline{\psfig{file=Hovatta_f3.eps,width=\textwidth}}
\vspace*{8pt}
\caption{Variability index of fractional polarization of LAT-detected (shaded) and non-LAT-detected (unshaded) sources from 2002 until August 2008 (left panel) and variability index of EVPA (right panel).\label{f3}}
\end{figure}
\section{Discussion}
By comparing the median fractional polarization since August 2008 with median values between 2002 and 2008,
we find the sources to be in a higher polarization state when detected by the LAT.
This complements our previous findings that the LAT-detected sources have higher apparent jet speeds,\cite{lister09b}
higher core brightness temperatures,\cite{kovalev09} larger apparent opening angles\cite{pushkarev09} and larger
Doppler boosting factors,\cite{savolainen09} In Kovalev et al.\cite{kovalev09} it was also shown that the LAT-detected
sources are preferentially found in higher radio activity states during the LAT-era.
The differences seen between detected and non-detected objects can be due to Doppler-beaming.\cite{lister05,lister09b,kovalev09} As is shown in Savolainen et al.\cite{savolainen09},
the LAT-detected sources are more boosted, which enhances their observed luminosity more than their non-detected less-beamed counterparts.
A higher polarization of LAT-detected sources may be explained if the magnetic field in the core is more ordered during the radio / $\gamma$-ray
flares. Indications that the intrinsic de-aberrated viewing angle in which the source is seen plays a significant role in whether a
jet is detected at $\gamma$-rays or not, are also seen.\cite{savolainen09} If the radio variations are caused by transverse shocks,\cite{hughes85} the observed fractional polarization can be modeled using the jet rest-frame viewing angle.\cite{hughes85,wardle94}
By assuming that the 15 GHz core is a $\tau = 1$
surface and using the jet rest-frame viewing angles from Ref.~\refcite{savolainen09}, we estimated the observed fractional polarization
for different shock strengths and fractions of uniform magnetic field using simulated values of maximum fractional polarization from
Ref.~\refcite{homan09} (their Figure A1) and equations of Ref.~\refcite{wardle94}. A uniform shock strength
value for all sources does not reproduce our observations very well. If we let the shock strength vary and choose cases which reproduce the
observed fractional polarization within 1\%, we see indications that the shocks in BL Lacertae objects
are stronger than in flat spectrum radio quasars. Further studies are required to confirm our results.
By using the EGRET observations, Lister et al.\cite{lister05} found no differences in the 15 GHz core polarization properties of $\gamma$-ray-detected and non-detected objects. They only used single-epoch VLBA data while we have used quasi-simultaneous LAT and VLBA
observations. However, the number of LAT-detected sources in our sample is small and therefore we intend to verify our results
using the LAT 1-year catalog. In this future study we will also examine the polarization of jet components located downstream from the core.
\section*{Acknowledgments}
The authors wish to acknowledge the contributions of the rest of the MOJAVE team. The MOJAVE project is supported under National
Science Foundation grant AST-0807860 and NASA Fermi grant NNX08AV67G. T.~H. wishes to acknowledge Magnus
Ehrnrooth Stiftelse for travel funds. Y.~Y.~K. is supported in part by the Russian Foundation for Basic Research
(project 08-02-00545) and a return fellowship of the Alexander von
Humboldt Foundation. T.~S.\ is a research fellow of the Alexander von~Humboldt Foundation. T.~S.\ also acknowledges a support by the Academy of Finland grant 120516. The VLBA is a facility of the National Science Foundation operated by the National
Radio Astronomy Observatory under cooperative agreement with Associated Universities, Inc.
|
\section{Introduction}\label{s:intro}
Let $M$ be a Calabi-Yau (CY) manifold. Roughly speaking, a submanifold
$L\subset M$ is \textit{special Lagrangian} (SL) if it is both minimal and
Lagrangian with respect to the ambient Riemannian and symplectic structures.
From the point of view of Riemannian Geometry it is of course natural to focus
on the minimality condition. It turns out that SLs are automatically
volume-minimizing in their homology class. In fact, this was Harvey and Lawson's
main motivation for defining and studying SLs within the general context of
Calibrated Geometry \cite{harveylawson}. This is still the most common point of
view on SLs and leads to emphasizing the role of analytic and Geometric Measure
Theory techniques. It also provides a connection with various classical problems
in Analysis such as the Plateau problem and the study of area-minimizing cones.
In many ways it is the point of view adopted here.
From the point of view of Symplectic Geometry it is instead natural to focus on
the Lagrangian condition. Specifically, SLs are examples of Maslov-zero
Lagrangian submanifolds. This leads to emphasizing the role of Symplectic
Topology techniques, both classical (such as the h-principle and moment maps)
and contemporary (such as Floer homology). An early instance of this point of
view is the work of Audin \cite{audin}; it also permeates the paper
\cite{haskinspacini} by Haskins and the author.
Given this richness of ingredients it is perhaps not surprising that SLs are
conjectured to play an important role in Mirror Symmetry \cite{kontsevich},
\cite{syz} and to produce interesting new invariants of CY manifolds
\cite{joyce:3spheres}. Likewise, and more intrinsically, they also tend to
exhibit other nice technical features. In particular it is by now well
understood that SLs often generate smooth, finite-dimensional, moduli spaces.
This SL deformation problem has been studied by a number of authors under
various topological and geometric assumptions. One clear path is the chain of
results initiated by McLean \cite{mclean}, who studied deformations of smooth
compact SLs; continued by the author \cite{pacini:defs} and Marshall
\cite{marshall}, who adapted that set-up to study certain smooth non-compact
(\textit{asymptotically conical}, AC) SLs; and further advanced by Joyce, who
presented analogous results for compact \textit{conically singular} (CS) SLs
\cite{joyce:II}.
The above three classes of SLs are intimately linked, as follows. One of the main open questions in SL geometry is how to compactify McLean's moduli spaces. This problem is currently one of the biggest obstructions to progress on the above conjectures. Roughly speaking, compactifying the moduli space requires adding to it a ``boundary'' containing singular compact SLs. By definition, CS SLs have isolated singularities modelled on SL cones in $\C^m$: they would be the simplest objects appearing in this boundary. If a CS SL appears in the boundary, it must be a limit of a 1-parameter family of smooth compact SLs. These smooth SLs can be recovered via a gluing construction which desingularizes the CS SL: (i) each singularity of the CS SL defines a SL cone in $\C^m$; (ii) each of these cones must admit a 1-parameter family of SL desingularizations, \textit{i.e.} AC SLs in $\C^m$ converging to the cone as the parameter $t$ tends to $0$; (iii) the family of smooth SLs is obtained by gluing the AC SLs into a neighbourhood of the singularities of the CS SL. This picture is made precise by Joyce's gluing results \cite{joyce:III}, \cite{joyce:IV},
\cite{joyce:V}. Section 8 of \cite{joyce:V} then shows that, in some cases and near the boundary,
the compactified moduli space can be locally written as a product of moduli
spaces of AC and CS SLs.
The above classes of submanifolds are special cases within the broader
category of \textit{Riemannian conifolds}, which includes manifolds
exhibiting both AC and CS ends. In other words, it allows CS SLs to become non-compact by allowing the presence of AC ends. This is of fundamental importance for the construction of SLs in $\C^m$: it is well-known that $\C^m$ does not admit any compact (smooth or singular) volume-minimizing submanifolds. Cones in $\C^m$ with an isolated singularity at
the origin are the simplest example of conifold: the construction of new examples and the study of their properties is currently one of the most active areas of SL research \cite{harveylawson},
\cite{haskins}, \cite{haskinskapouleas}, \cite{haskinspacini}, \cite{joyce:symmetries}, \cite{ohnita}. Conifolds provide the appropriate framework in which to extend all the above research. In particular, they might also substitute AC SLs in Joyce's gluing results: one could try to cut out a conical singularity of the CS SL and replace it with a different singular conifold, thus jumping from one area of the boundary of the compactified moduli space, containing certain CS SLs, to another.
The paper at hand is Part I of a multi-step project aiming to set up a general theory of SL conifolds and, more generally, of calibrated conifolds. Two other papers related to this project are currently available: \cite{pacini:weighted}, \cite{pacini:slgluing}. Further work is in progress. The goal of this paper is to provide a general deformation theory of SL conifolds in $\C^m$. The best set-up for the SL deformation
problem is the one provided by Joyce \cite{joyce:II}. It is based on his
Lagrangian neighbourhood and regularity theorems \cite{joyce:I}. Joyce's
framework has two benefits: (i) it simplifies the Analysis via a reduction from
the semi-elliptic operator $d\oplus d^*$ on 1-forms to the
elliptic Laplace operator on functions, (ii) it nicely emphasizes the separate
contributions to the dimension of $\mathcal{M}_L$ coming from the topological
and from the analytic components. Along with the main result Theorem
\ref{thm:accssl} concerning moduli spaces of CS/AC SL submanifolds in $\C^m$, we thus
present new proofs of the previously-known results, emphasizing this point of view.
In this sense, this paper also serves the purpose of surveying and unifying those results. More importantly, it lays down the geometric foundations for \cite{pacini:slgluing}; the analytic foundations are provided by \cite{pacini:weighted}.
We now summarize the contents of this paper. Section \ref{s:geometry_review}
introduces the category of $m$-dimensional Riemannian conifolds. The main
definitions are standard but Section \ref{ss:closedforms} contains an
investigation into the structure of various spaces of closed 1-forms on these
manifolds. This is a fundamental component of the Lagrangian and SL deformation
theory. The corresponding notion of ``subconifolds'' is presented in Section
\ref{s:lagconifolds}, leading to the concept of \textit{Lagrangian conifolds}.
Deformation theory begins in Section \ref{s:lagdefs}. From various points of
view it seems most satisfying to begin with the general (infinite-dimensional)
theory of Lagrangian deformations. This is presented as a direct consequence of
Joyce's Lagrangian neighbourhood theorems, coupled with the material of Section
\ref{ss:closedforms}. The case of Lagrangian cones is studied in particular
detail in Section \ref{ss:conelagdefs} as it provides the backbone for all other
cases. After presenting the necessary definitions in Section \ref{s:slgeometry},
the analogous framework for deforming SL conifolds is developed in Section
\ref{s:setup}. With the aim of making this paper reasonably self-contained,
Section \ref{s:reviewlaplace} summarizes from \cite{pacini:weighted} some
results concerning harmonic functions on conifolds. The SL deformation theory is
then completed in Section \ref{s:moduli}. The proofs rely upon a fair amount of
analytic machinery: weighted Sobolev spaces, embedding theorems and the theory
of elliptic operators on conifolds. Full details are provided in \cite{pacini:weighted}.
\
\textbf{Important remark:} To simplify certain arguments, throughout this paper
we assume $m\geq 3$.
\section{Geometry of conifolds}\label{s:geometry_review}
\subsection{Asymptotically conical and conically singular
manifolds}\label{ss:accs_review}
We introduce here the categories of differentiable and Riemannian manifolds
mainly relevant to this paper, referring to \cite{pacini:weighted} for further
details. Following \cite{joyce:I}, however, we introduce a small variation of
the notion of ``conically singular" manifolds: presenting them in terms of the
compactification $\bar{L}$ will allow us to keep track of the singular points
$x_i$. This plays no role in this section but in Section \ref{s:lagdefs} it will
become very useful.
\begin{definition}\label{def:manifold_ends}
Let $L^m$ be a smooth manifold. We say $L$ is a \textit{manifold with ends} if
it satisfies the following conditions:
\begin{enumerate}
\item We are given a compact subset $K\subset L$ such that $S:=L\setminus K$ has
a finite number of connected components $S_1,\dots,S_e$, \textit{i.e.}
$S=\amalg_{i=1}^e S_i$.
\item For each $S_i$ we are given a connected ($m-1$)-dimensional compact
manifold $\Sigma_i$ without boundary.
\item There exist diffeomorphisms $\phi_i:\Sigma_i\times [1,\infty)\rightarrow
\overline{S_i}$.
\end{enumerate}
We then call the components $S_i$ the \textit{ends} of $L$ and the manifolds
$\Sigma_i$ the \textit{links} of $L$. We denote by $S$ the union of the ends and
by $\Sigma$ the union of the links of $L$.
\end{definition}
\begin{definition}\label{def:metrics_ends}
Let L be a manifold with ends. Let $g$ be a Riemannian metric on $L$. Choose an
end $S_i$ with corresponding link $\Sigma_i$.
We say that $S_i$ is a \textit{conically singular} (CS) end if the following
conditions hold:
\begin{enumerate}
\item $\Sigma_i$ is endowed with a Riemannian metric $g_i'$.
We then let $(\theta,r)$ denote the generic point on the product manifold
$C_i:=\Sigma_i\times (0,\infty)$ and $\tg_i:=dr^2+r^2g_i'$ denote the
corresponding \textit{conical metric} on $C_i$.
\item There exist a constant $\nu_i>0$ and a diffeomorphism
$\phi_i:\Sigma_i\times (0,\epsilon]\rightarrow \overline{S_i}$ such that, as $r\rightarrow
0$ and for all $k\geq 0$,
$$|\tnabla^k(\phi_i^*g-\tg_i)|_{\tg_i}=O(r^{\nu_i-k}),$$
where $\tnabla$ is the Levi-Civita connection on $C_i$ defined by $\tg_i$.
\end{enumerate}
We say that $S_i$ is an \textit{asymptotically conical} (AC) end if the
following conditions hold:
\begin{enumerate}
\item $\Sigma_i$ is endowed with a Riemannian metric $g_i'$.
We again let $(\theta,r)$ denote the generic point on the product manifold
$C_i:=\Sigma_i\times (0,\infty)$ and $\tg_i:=dr^2+r^2g_i'$ denote the
corresponding conical metric on $C_i$.
\item There exist a constant $\nu_i<0$ and a diffeomorphism
$\phi_i:\Sigma_i\times [R,\infty)\rightarrow \overline{S_i}$ such that, as $r\rightarrow
\infty$ and for all $k\geq 0$,
$$|\tnabla^k(\phi_i^*g-\tg_i)|_{\tg_i}=O(r^{\nu_i-k}),$$
where $\tnabla$ is the Levi-Civita connection on $C_i$ defined by $\tg_i$.
\end{enumerate}
In either of the above situations we call $\nu_i$ the \textit{convergence rate}
of $S_i$.
\end{definition}
We refer to \cite{pacini:weighted} for a better understanding of the asymptotic
conditions introduced in Definition \ref{def:metrics_ends}.
\begin{definition}\label{def:cs_manifold}
Let $(\bar{L},d)$ be a metric space. $\bar{L}$ is a \textit{Riemannian manifold
with conical singularities} (CS manifold) if it satisfies the following
conditions.
\begin{enumerate}
\item We are given a finite number of points $\{x_1,\dots,x_e\}\in \bar{L}$ such that $L:=\bar{L}\setminus\{x_1,\dots,x_e\}$ has the
structure of a smooth $m$-dimensional manifold with $e$ ends.
More specifically, we assume given $\epsilon\in (0,1)$ such that any pair of
distinct points satisfies $d(x_i,x_j)>2\epsilon$. Set $S_i:=\{x\in L:
0<d(x,x_i)<\epsilon\}$. We then assume that $S_i$ are the ends of $L$ with
respect to some given connected links $\Sigma_i$.
\item We are given a Riemannian metric $g$ on $L$ inducing the distance $d$.
\item With respect to $g$, each end $S_i$ is CS in the sense of Definition
\ref{def:metrics_ends}.
\end{enumerate}
It follows from our definition that any CS manifold $\bar{L}$ is compact. We
will often not distinguish between $\bar{L}$ and $L$, but notice that $(L,g)$ is
neither compact nor complete. We call $x_i$ the \textit{singularities} of
$\bar{L}$.
\end{definition}
\begin{definition}\label{def:ac_manifold}
Let $(L,g)$ be a Riemannian manifold. $L$ is a \textit{Riemannian manifold with
asymptotically conical ends} (AC manifold) if it satisfies the following
conditions.
\begin{enumerate}
\item $L$ is a smooth manifold with $e$ ends $S_i$ and connected links
$\Sigma_i$.
\item Each end $S_i$ is AC in the sense of Definition \ref{def:metrics_ends}.
\end{enumerate}
\end{definition}
One can check that AC manifolds are non-compact but complete.
\begin{definition} \label{def:accs_manifold}
Let $(\bar{L},d)$ be a metric space. We say that $\bar{L}$ is a
\textit{Riemannian CS/AC manifold} if it satisfies the following conditions.
\begin{enumerate}
\item We are given a finite number of points $\{x_1,\ldots,x_s\}$ and a number $l$ such that
$L:=\bar{L}\setminus\{x_1,\dots,x_s\}$ has the structure of a smooth
$m$-dimensional manifold with $s+l$ ends.
\item We are given a metric $g$ on $L$ inducing the distance $d$.
\item With respect to $g$, neighbourhoods of the points $x_i$ have the structure
of CS ends in the sense of Definition \ref{def:metrics_ends}. These are the
``small" ends. We also assume that the remaining ends are ``large",
\textit{i.e.} they have the structure of AC ends in the sense of Definition
\ref{def:metrics_ends}.
\end{enumerate}
We will denote the union of the CS links (respectively, of the CS ends) by
$\Sigma_0$ (respectively, $S_0$) and those corresponding to the AC links and
ends by $\Sigma_\infty$, $S_\infty$.
\end{definition}
\begin{definition} \label{def:conifold}
We use the generic term \textit{conifold} to indicate any CS, AC or CS/AC
manifold. If $(L,g)$ is a conifold and $C:=\amalg C_i$ is the union of the
corresponding cones as in Definition \ref{def:metrics_ends}, endowed with the
induced metric $\tg$, we say that $(L,g)$ is \textit{asymptotic} to $(C,\tg)$.
\end{definition}
\begin{remark}\label{rem:nondistinct_sings}
If we think of $\bar{L}$ as a generic compactification of the manifold with ends $L$, we should allow several CS ends to become connected by the addition of a single singular point. Notice however that we have imposed that our links be connected. We should thus allow that our points $x_i$ be not necessarily distinct. This apparent detail becomes extremely relevant when working with ``parametric connect sums'', as in \cite{pacini:weighted}, \cite{pacini:slgluing}. In \cite{pacini:weighted}, however, we do not need to mention it because there the connect sum $L_t$ is defined in terms of $L$: in some sense, the compactification $\bar{L}$ appears only \textit{a posteriori} with respect to the connect sum, as the limit of $L_t$ as $t\rightarrow 0$. In \cite{pacini:slgluing} we again do not need to mention it, this time because the connect sum is defined in terms of an immersion: by definition, the immersion is allowed to identify points so we might as well assume that the $x_i$ and cones are initially distinct. The connect sum then depends only on the identifications determined by the immersion.
\end{remark}
Cones in $\R^n$ are of course the archetype of CS/AC manifold, as follows.
\begin{definition} \label{def:cone}
A subset $\bar{\mathcal{C}}\subseteq\R^n$ is a \textit{cone} if it is invariant
under dilations of $\R^n$, \textit{i.e.} if $t\cdot \bar{\mathcal{C}}\subseteq
\bar{\mathcal{C}}$, for all $t\geq 0$. It is uniquely identified by its
\textit{link} $\Sigma:=\bar{\mathcal{C}}\bigcap \Sph^{n-1}$. We will set
$\mathcal{C}:=\bar{\mathcal{C}}\setminus 0$. The cone is \textit{regular} if
$\Sigma$ is smooth. From now on we will always assume this.
Let $g'$ denote the induced metric on $\Sigma$. Then $\mathcal{C}$ with its
induced metric is isometric to $\Sigma\times (0,\infty)$ with the conical metric
$\tg:=dr^2+r^2g'$. In particular $\bar{\mathcal{C}}$ is a CS/AC manifold; it has
as many AC and CS ends as the number of connected components $\Sigma_i$ of
$\Sigma$. Each $\Sigma_i$ thus defines a singular point $x_i$ but these singular
points are not distinct: they all coincide with the origin. Notice that $\Sigma$
is a subsphere $\Sph^{m-1}\subseteq \Sph^{n-1}$ iff $\bar{\mathcal{C}}$ is an
$m$-plane in $\R^n$.
\end{definition}
Let $E$ be a vector bundle over $(L,g)$. Assume $E$ is endowed with a metric and
metric connection $\nabla$: we say that $(E,\nabla)$ is a \textit{metric pair}.
In later sections $E$ will usually be a bundle of differential forms $\Lambda^r$
on $L$, endowed with the metric and Levi-Civita connection induced from $g$. We
can define two types of Banach spaces of sections of $E$,
referring to \cite{pacini:weighted} for further details regarding the structure
and properties of these spaces.
Regarding notation, given a vector
$\boldsymbol{\beta}=(\beta_1,\dots,\beta_e)\in \R^e$ and $j\in\N$ we set
$\boldsymbol{\beta}+j:=(\beta_1+j,\dots,\beta_e+j)$. We write
$\boldsymbol{\beta}\geq\boldsymbol{\beta}'$ iff $\beta_i\geq\beta_i'$.
\begin{definition}\label{def:csac_sectionspaces}
Let $(L,g)$ be a conifold with $e$ ends. We say that a smooth function
$\rho:L\rightarrow (0,\infty)$ is a \textit{radius function} if $\rho(x)\equiv
r$ on each end. Given any vector
$\boldsymbol{\beta}=(\beta_1,\dots,\beta_{e})\in\R^{e}$, choose a function
$\boldsymbol{\beta}$ on $L$ which, on each end $S_i$, restricts to the constant
$\beta_i$.
Given any metric pair $(E,\nabla)$, the \textit{weighted Sobolev spaces} are
defined by
\begin{equation}\label{eq:weighted_sob}
W^p_{k;\boldsymbol{\beta}}(E):=\mbox{Banach space completion of the space
}\{\sigma\in C^\infty(E):\|\sigma\|_{W^p_{k;\boldsymbol{\beta}}}<\infty\},
\end{equation}
where we use the norm
$\|\sigma\|_{W^p_{k;\boldsymbol{\beta}}}:=(\Sigma_{j=0}^k\int_L|\rho^{
-\boldsymbol{\beta}+j}\nabla^j\sigma|^p\rho^{-m}\,\mbox{vol}_g)^{1/p}$.
The \textit{weighted spaces of $C^k$ sections} are defined by
\begin{equation}\label{eq:weighted_C^k}
C^k_{\boldsymbol{\beta}}(E):=\{\sigma\in C^k(E):
\|\sigma\|_{C^k_{\boldsymbol{\beta}}}<\infty\},
\end{equation}
where we use the norm $\|\sigma\|_{C^k_{\boldsymbol{\beta}}}:=\sum_{j=0}^k
\mbox{sup}_{x\in L}|\rho^{-\boldsymbol{\beta}+j}\nabla^j\sigma|$. Equivalently,
$C^k_{\boldsymbol{\beta}}(E)$ is the space of sections $\sigma\in C^k(E)$ such
that $|\nabla^j \sigma|=O(r^{\boldsymbol{\beta}-j})$ as $r\rightarrow 0$
(respectively, $r\rightarrow\infty$) along each CS (respectively, AC) end. These
are also Banach spaces.
To conclude, the \textit{weighted space of smooth sections} is defined by
\begin{equation*}
C^\infty_{\boldsymbol{\beta}}(E):=\bigcap_{k\geq 0} C^k_{\boldsymbol{\beta}}(E).
\end{equation*}
Equivalently, this is the space of smooth sections such that $|\nabla^j
f|=O(\rho^{\boldsymbol{\beta}-j})$ for all $j\geq 0$. This space has a natural
Fr\'echet structure.
When $E$ is the trivial $\R$ bundle over $L$ we obtain weighted spaces of
functions on $L$. We usually denote these by $W^p_{k,\boldsymbol{\beta}}(L)$ and
$C^k_{\boldsymbol{\beta}}(L)$. In the case of a CS/AC manifold we will often
separate the CS and AC weights, writing
$\boldsymbol{\beta}=(\boldsymbol{\mu},\boldsymbol{\lambda})$ for some
$\boldsymbol{\mu}\in \R^s$ and some $\boldsymbol{\lambda}\in \R^l$. We then
write $C^k_{(\boldsymbol{\mu},\boldsymbol{\lambda})}(E)$ and
$W^p_{k,(\boldsymbol{\mu},\boldsymbol{\lambda})}(E)$.
\end{definition}
For these spaces one can prove the validity of the following weighted version of
the Sobolev Embedding Theorems, cf. \cite{pacini:weighted}.
\begin{theorem}\label{thm:embedding}
Let $(L,g)$ be an AC manifold. Let $(E,\nabla)$ be a metric pair over $L$.
Assume $k\geq 0$, $l\in\{1,2,\dots\}$ and $p\geq 1$. Set
$p^*_l:=\frac{mp}{m-lp}$. Then, for all
$\boldsymbol{\beta}'\geq\boldsymbol{\beta}$,
\begin{enumerate}
\item If $lp<m$ then there exists a continuous embedding
$W^p_{k+l,\boldsymbol{\beta}}(E)\hookrightarrow
W^{p^*_l}_{k,\boldsymbol{\beta}'}(E)$.
\item If $lp=m$ then, for all $q\in [p,\infty)$, there exist continuous
embeddings $W^p_{k+l,\boldsymbol{\beta}}(E)\hookrightarrow
W^q_{k,\boldsymbol{\beta}'}(E)$.
\item If $lp>m$ then there exists a continuous embedding
$W^p_{k+l,\boldsymbol{\beta}}(E)\hookrightarrow C^k_{\boldsymbol{\beta}'}(E)$.
\end{enumerate}
Furthermore, assume $lp>m$ and $k\geq 0$. Then the corresponding weighted
Sobolev spaces are closed under multiplication, in the following sense. For
any $\boldsymbol{\beta}_1$ and $\boldsymbol{\beta_2}$ there exists $C>0$ such
that, for all $u\in W^p_{k+l,\boldsymbol{\beta_1}}$ and $v\in
W^p_{k+l,\boldsymbol{\beta_2}}$,
\begin{equation*}
\|uv\|_{W^p_{k+l,\boldsymbol{\beta_1}+\boldsymbol{\beta_2}}}\leq
C\|u\|_{W^p_{k+l,\boldsymbol{\beta_1}}}\|v\|_{W^p_{k+l,\boldsymbol{\beta_2}}}.
\end{equation*}
Let $(L,g)$ be a CS manifold. Then the same conclusions hold for all
$\boldsymbol{\beta}'\leq\boldsymbol{\beta}$.
Let $(L,g)$ be a CS/AC manifold. Then, setting
$\boldsymbol{\beta}=(\boldsymbol{\mu},\boldsymbol{\lambda})$, the same
conclusions hold for $\boldsymbol{\mu}'\leq\boldsymbol{\mu}$ on the CS ends and
$\boldsymbol{\lambda}'\geq\boldsymbol{\lambda}$ on the AC ends.
\end{theorem}
\subsection{Weighted $1$-forms and cohomology}\label{ss:closedforms}
Any smooth compact manifold or smooth manifold with ends $L$ has topology of
finite type. In particular, the first cohomology group
$$H^1(L;\R):=\frac{\{\mbox{Smooth closed $1$-forms on $L$}\}}{d(C^\infty(L))}$$
has finite dimension $b^1(L)$, proving the following statement concerning the
structure of the space of smooth closed $1$-forms.
\begin{decomp}[for compact manifolds or manifolds with
ends]\label{decomp:closedforms}
Let $L$ be a smooth compact manifold or a smooth manifold with ends. Choose a
finite-dimensional vector space $H$ of closed $1$-forms on $L$ such that the map
\begin{equation}
H\rightarrow H^1(L;\R), \ \ \alpha\mapsto [\alpha]
\end{equation}
is an isomorphism. Then
\begin{equation}\label{eq:closedforms}
\{\mbox{Smooth closed $1$-forms on $L$}\}= H\oplus d(C^\infty(L)).
\end{equation}
\end{decomp}
We now want to show that in the case of a manifold with ends there exist natural
conditions on the space of 1-forms $H$.
\begin{definition}\label{def:translationinvariant}
Given a manifold $\Sigma$, set $C:=\Sigma\times (0,\infty)$. Consider the
projection $\pi:\Sigma\times (0,\infty)\rightarrow \Sigma$. A $p$-form $\eta$ on
$C$ is \textit{translation-invariant} if it is of the form $\eta=\pi^*\eta'$,
for some $p$-form $\eta'$ on $\Sigma$.
\end{definition}
\begin{lemma}\label{lemma:formclosed}
Let $L$ be a smooth manifold with ends $S_i$. Let $\alpha$ be a smooth closed
1-form on $L$. Then there exist a smooth closed 1-form $\alpha'$ and a smooth
function $A$ on $L$ such that $\alpha'_{|S_i}$ is translation-invariant and
$\alpha=\alpha'+dA$.
If furthermore $\alpha$ has compact support then we can choose $\alpha'$ to have
compact support.
\end{lemma}
\begin{proof}
The proof follows the scheme of the Poincar\'e Lemma for de Rham cohomology, cf.
\textit{e.g.} \cite{botttu}. Given any $p$-form $\eta$ on $S_i=\Sigma_i\times
(1,\infty)$, we can write
$$\eta=\eta_1(\theta,r)+\eta_2(\theta,r)\wedge dr$$
for some $r$-dependent $p$-form $\eta_1$ and ($p-1$)-form $\eta_2$ on $\Sigma$.
Specifically, $\eta_1$ is the restriction of $\eta$ to the cross-sections
$\Sigma_i\times\{r\}$ and $\eta_2:=i_{\partial r} \eta$. For a fixed $R_0>1$ we
then define $(K\eta)(\theta,r):=\int_{R_0}^r\eta_2(\theta,\rho)\,d\rho$.
Let us apply this to the 1-form obtained by restricting $\alpha$ to $S_i$,
writing
$$\alpha_{|S_i}=\alpha_1(\theta,r)+\alpha_2(\theta,r)\, dr$$
for some $r$-dependent 1-form $\alpha_1$ and function $\alpha_2$ on $\Sigma_i$.
It is then easy to check that
\begin{eqnarray*}
d\alpha_{|S_i} &=& d_\Sigma\alpha_1-(\frac{\partial}{\partial r} \alpha_1)\wedge
dr+(d_\Sigma\alpha_2)\wedge dr,\\
K\alpha_{|S_i} &=&\int_{R_0}^r\alpha_2(\theta,\rho)\,d\rho,\\
d(K\alpha_{|S_i})
&=&\int_{R_0}^rd_\Sigma\alpha_2(\theta,\rho)\,d\rho+\alpha_2(\theta,r)\,dr.
\end{eqnarray*}
From $d\alpha=0$ it follows that $\alpha_1(\theta,R_0)+d(K\alpha)=\alpha_{|S_i}$
and that $\alpha_1(\theta,R_0)$ is closed. Setting
$\alpha'_i:=\alpha_1(\theta,R_0)$ and $A_i:=K\alpha$ we can rewrite this as
$\alpha_{|S_i}=\alpha'_i+dA_i$. Interpolating between the $A_i$ yields a global
smooth function $A$ on $L$ such that $\alpha_{|S_i}=\alpha'_i+dA_{|S_i}$. We can
now define $\alpha':=\alpha-dA$ to obtain the global relationship
$$\alpha=\alpha'+dA.$$
It is clear from this construction that if $\alpha$ has compact support then
(choosing $R_0$ large enough) $\alpha'$ also has compact support.
\end{proof}
Recall that compactly-supported forms give rise to the following theory. Let $L$
be a smooth manifold with ends. We denote by $\Lambda^p_c(L;\R)$ the space of
smooth compactly-supported $p$-forms on $L$ and by $H^p_c(L;\R)$ the
corresponding cohomology groups. Let $\Sigma$ denote the union of the links of
$L$. Notice that $L$ is deformation-equivalent to a compact manifold with
boundary $\Sigma$. Standard algebraic topology (see also \cite{joyce:I} Section
2.4) proves that the inclusion $\Sigma\subset L$ gives rise to a long exact
sequence in cohomology
\begin{equation}\label{eq:cohomsequence}
0\rightarrow H^0(L;\R)\rightarrow H^0(\Sigma;\R)\stackrel{\delta}{\rightarrow}
H^1_c(L;\R)\stackrel{\gamma}{\rightarrow} H^1(L;\R)\stackrel{\rho}{\rightarrow}
H^1(\Sigma;\R)\rightarrow\dots.
\end{equation}
Here, $\gamma$ is induced by the injection $\Lambda^1_c(L;\R) \rightarrow
\Lambda^1(L;\R)$ and $\rho$ is induced by the restriction
$\Lambda^1(L;\R)\rightarrow \Lambda^1(\Sigma_0;\R)$. We set
$\widetilde{H}^1_c:=\mbox{Im}(\gamma)=\mbox{Ker}(\rho)$. Exactness implies that
\begin{align}\label{eq:dimexactseq}
\mbox{dim}(\widetilde{H}^1_c) &=
\mbox{dim}(H^1_c(L;\R))-\mbox{dim}(H^0(\Sigma;\R))+\mbox{dim}(H^0(L;\R))\\
\nonumber
&= b^1_c(L)-e+1.
\end{align}
\begin{remark} \label{rem:compactsupport}
The sequence \ref{eq:cohomsequence} shows that
\begin{equation}\label{eq:exactseq}
H^1_c(L,\R)\simeq \widetilde{H}^1_c\oplus \mbox{Ker}(\gamma) = \widetilde{H}^1_c
\oplus \mbox{Im}(\delta).
\end{equation}
This decomposition can be expressed in words as follows. By definition,
$H^1_c(L;\R)$ is determined by the classes of compactly-supported 1-forms which
are not the differential of a compactly-supported function. Given any such form,
there are two cases: (i) it is not the differential of \textit{any} function, in
which case $\gamma$ maps its class to a non-zero element of $\widetilde{H}^1_c$,
(ii) it is the differential of some function, in which case $\gamma$ maps its
class to zero. However, this function is necessarily constant on the ends of
$L$: these constants can be parametrized via $H^0(\Sigma;\R)$. Notice that the
function is only well-defined up to a constant; likewise, $\mbox{Im}(\delta)$
coincides with $H^0(\Sigma;\R)$ only up to $H^0(L;\R)\simeq\R$.
\end{remark}
We can now choose $H$ as follows. For $i=1,\dots,k=\mbox{dim}(\tilde{H}^1_c)$
let $[\alpha_i]$ be a basis of $\widetilde{H}^1_c$. According to Lemma
\ref{lemma:formclosed} we can choose $\alpha_i'$ with compact support such that
[$\alpha_i']=[\alpha_i]$. For $i=1,\dots,N=\mbox{dim}(H^1)$ let $[\alpha_i]$
denote an extension to a basis of $H^1(L;\R)$. Again using Lemma
\ref{lemma:formclosed} we can choose an extension $\alpha_i'$ of
translation-invariant 1-forms such that [$\alpha_i']=[\alpha_i]$. Set
\begin{equation}\label{eq:naturalH}
\widetilde{H}:=\mbox{span}\{\alpha_1',\dots,\alpha_k'\},\ \
H:=\mbox{span}\{\alpha_1',\dots,\alpha_N'\}.
\end{equation}
Then $H$ satisfies the assumptions of Decomposition \ref{decomp:closedforms}.
One advantage of this choice of $H$ is that it reflects the relationship of
$\widetilde{H}^1_c$ to $H^1$. Specifically, if we apply Decomposition
\ref{decomp:closedforms} to $\alpha$ writing $\alpha=\alpha'+dA$ with
$\alpha'\in H$, then $[\alpha]\in \widetilde{H}^1_c$ iff $\alpha'\in
\widetilde{H}$, \textit{i.e.} iff $\alpha'$ has compact support.
We now want to achieve analogous decompositions for CS and AC manifolds, in
terms of weighted spaces of closed and exact $1$-forms.
\begin{lemma}\label{l:translationinvariant}
Let $(\Sigma,g')$ be a Riemannian manifold. Let the corresponding cone $C$ have
the conical metric $\tg:=dr^2+r^2g'$. Then any translation-invariant $p$-form
$\eta=\pi^*\eta'$ belongs to the weighted space $C^\infty_{(-p,-p)}(\Lambda^p)$.
For any $\beta>0$, $\eta$ belongs to the smaller weighted space
$C^\infty_{(-p+\beta,-p-\beta)}(\Lambda^p)$ iff $\eta'=0$.
\end{lemma}
\begin{proof} As seen in the proof of Lemma \ref{lemma:formclosed}, the general
$p$-form $\eta$ on $C$ can be written
$\eta=\eta_1(\theta,r)+\eta_2(\theta,r)\wedge dr$. The form is
translation-invariant iff $\eta_1$ is $r$-independent and $\eta_2=0$. In this
case $|\eta|_{\tg}=r^{-p}|\eta_1|_{g'}$ so $|\eta|_{\tg}=O(r^{-p})$ both for
$r\rightarrow 0$ and for $r\rightarrow \infty$. This proves that $\eta\in
C^0_{(-p,-p)}(\Lambda^p)$. To show that $\eta\in C^\infty_{(-p,-p)}(\Lambda^p)$
it is necessary to estimate $|\tnabla^k\eta|_{\tg}$, where $\tnabla$ is the
Levi-Civita connection. This can be done fairly explicitly in terms of
Christoffel symbols. In particular one can choose local coordinates on
$U\subset\Sigma$ defining a local frame $\partial_1,\cdots,\partial_{m-1}$. Set
$\partial_0:=\partial r$, the standard frame on $(0,\infty)$. The Christoffel
symbols for the corresponding frame on $(0,\infty)\times U$ and the metric $\tg$
can then be computed explicitly: for $i,j,k\geq 1$ one finds that
$\tGamma_{i,j}^k$ is bounded, $\tGamma_{i,j}^0=O(r)$,
$\tGamma_{i,0}^k=O(r^{-1})$,
$\tGamma_{0,0}^k=\tGamma_{i,0}^0=\tGamma_{0,0}^0=0$. The Christoffel symbols
defined by $\tg$ for the other tensor bundles depend linearly on these, so they
have the same bounds. Using these calculations one finds that
$|\tnabla^k\eta|_{\tg}=O(r^{-p-k})$, as desired.
It is clear from the proof that $\eta$ satisfies stronger bounds iff it
vanishes.
\end{proof}
\begin{decomp}[for CS or AC manifolds and forms with allowable
growth]\label{decomp:closedforms_growth}
Let $L$ be a CS manifold. Choose a finite-dimensional vector space $H$ of smooth
closed 1-forms on $L$ as in Equation \ref{eq:naturalH}. Then, for any
$\boldsymbol{\beta}<0$,
\begin{equation}\label{eq:closedforms_growth_cs}
\{\mbox{Closed 1-forms on $L$ in $C^\infty_{\boldsymbol{\beta}-1}(\Lambda^1)$}\}
= H\oplus d(C^\infty_{\boldsymbol{\beta}}(L)).
\end{equation}
Analogously, let $L$ be an AC manifold. Choose $H$ as above. Then, for any
$\boldsymbol{\beta}>0$,
\begin{equation}\label{eq:closedforms_growth_ac}
\{\mbox{Closed 1-forms on $L$ in $C^\infty_{\boldsymbol{\beta}-1}(\Lambda^1)$}\}
= H\oplus d(C^\infty_{\boldsymbol{\beta}}(L)).
\end{equation}
\end{decomp}
\begin{proof}
Consider the CS case. Since $\boldsymbol{\beta}<0$, Lemma
\ref{l:translationinvariant} proves that $H\oplus
d(C^\infty_{\boldsymbol{\beta}}(L))\subseteq \{\mbox{Closed 1-forms in
$C^\infty_{\boldsymbol{\beta}-1}(\Lambda^1)$}\}$. Now choose a closed $\alpha\in
C^\infty_{\boldsymbol{\beta}-1}(\Lambda^1)$. By Decomposition
\ref{decomp:closedforms} we can write $\alpha=\alpha'+dA$, for some $\alpha'\in
H$ and $A\in C^\infty(L)$. Notice that $dA=\alpha-\alpha'\in
C^\infty_{\boldsymbol{\beta}-1}(\Lambda^1)$. By integration, again using the
fact $\boldsymbol{\beta}<0$, we conclude that $A\in
C^\infty_{\boldsymbol{\beta}}(L)$. This proves the opposite inclusion, thus the
identity. The AC case is analogous.
\end{proof}
\begin{lemma}\label{l:formexact}
Assume $L$ is a CS manifold. If $\alpha$ is a smooth closed 1-form on $L$
belonging to the space $C^\infty_{\boldsymbol{\beta}-1}(\Lambda^1)$ for some
$\boldsymbol{\beta}>0$ then there exists a smooth closed 1-form $\alpha'$ with
compact support on $L$ and a smooth function $A\in
C^\infty_{\boldsymbol{\beta}}(L)$ such that $\alpha=\alpha'+dA$.
Assume $L$ is an AC manifold. If $\alpha$ is a smooth closed 1-form on $L$
belonging to the space $C^\infty_{\boldsymbol{\beta}-1}(\Lambda^1)$ for some
$\boldsymbol{\beta}<0$ then there exists a smooth closed 1-form $\alpha'$ with
compact support on $L$ and a smooth function $A\in
C^\infty_{\boldsymbol{\beta}}(L)$ such that $\alpha=\alpha'+dA$.
\end{lemma}
\begin{proof} The proof is a variation of the proof of Lemma
\ref{lemma:formclosed}, as follows. Consider the AC case. Write
$\alpha_{|S_i}=\alpha_1+\alpha_2\wedge dr$. Define
$K\alpha:=-\int_r^\infty\alpha_2(\theta,\rho)\,d\rho$: this converges because
$\boldsymbol{\beta}<0$. It is simple to check that $d(K\alpha)=\alpha$; in
particular, this shows that $\alpha$ is exact on each end $S_i$. Setting
$A:=K\alpha$ and extending as in Lemma \ref{lemma:formclosed} leads to a global
decomposition $\alpha=\alpha'+dA$ on $L$. By construction $\alpha'$ has compact
support and $A\in C^\infty_{\boldsymbol{\beta}}$. The CS case is analogous, with
$K\alpha:=\int_0^r\alpha_2(\theta,\rho)\,d\rho$.
\end{proof}
\begin{decomp}[for CS or AC manifolds and forms with allowable
decay]\label{decomp:closedforms_decay}
Let $L$ be a CS manifold. Assume $\boldsymbol{\beta}>0$. Choose a
finite-dimensional vector space $H$ of closed 1-forms on $L$ as in Equation
\ref{eq:naturalH}, using $\widetilde{H}_0$ to denote the space $\widetilde{H}$.
For any $i=1,\dots,e$ choose a smooth function $f_i$ on $L$ such that $f_i\equiv
1$ on the end $S_i$ and $f_i\equiv 0$ on the other ends. We can do this in such
a way that $\sum f_i\equiv 1$. Let $E_0$ denote the $e$-dimensional vector space
generated by these functions. By construction $E_0$ contains the constant
functions so $d(E_0)$ has dimension $e-1$. It is simple to check that
$d(E_0)\cap d(C^\infty_{\boldsymbol{\beta}}(L))=\{0\}$. Then
\begin{equation}\label{eq:closedforms_decay_cs}
\{\mbox{Closed 1-forms on $L$ in
$C^\infty_{\boldsymbol{\beta}-1}(\Lambda^1)$}\}=\widetilde{H}_0\oplus
d(E_0)\oplus d(C^\infty_{\boldsymbol{\beta}}(L)).
\end{equation}
Analogously, let $L$ be an AC manifold. Assume $\boldsymbol{\beta}<0$. Choose
spaces as above, this time using the notation $\widetilde{H}_\infty$ and
$E_\infty$. Then
\begin{equation}\label{eq:closedforms_decay_ac}
\{\mbox{Closed 1-forms on $L$ in $C^\infty_{\boldsymbol{\beta}-1}(\Lambda^1)$}\}
= \widetilde{H}_\infty\oplus d(E_\infty)\oplus
d(C^\infty_{\boldsymbol{\beta}}(L)).
\end{equation}
\end{decomp}
\begin{proof} Consider the CS case. The inclusion $\supseteq$ is clear.
Conversely, let $\alpha\in C^\infty_{\boldsymbol{\beta}-1}(\Lambda^1)$ be
closed. Decomposition \ref{decomp:closedforms} allows us to write
$\alpha=\alpha'+dA$, for some uniquely defined $\alpha'\in H$ and some $A\in
C^\infty(L)$, well-defined up to a constant. Lemma \ref{l:formexact} implies
that the cohomology class of $\alpha$ belongs to the space $\widetilde{H}^1_c$,
\textit{i.e.} that $\alpha'\in \widetilde{H}_0$ so it has compact support. This
shows that $dA\in C^\infty_{\boldsymbol{\beta}-1}(\Lambda^1)$. Writing
$A_i:=A_{|S_i}$ we find $dA_i=d_{\Sigma_i}A_i+\frac{\partial A_i}{\partial
r}\,dr$, thus $\frac{\partial A_i}{\partial r}\in
C^\infty_{\boldsymbol{\beta}-1}(L)$. This shows that $\int_0^r\frac{\partial
A_i}{\partial r}\,d\rho\in C^\infty_{\boldsymbol{\beta}}(L)$. This determines
$A_i$ up to a constant $c_i$ on each end. Together with Equation
\ref{eq:exactseq} this proves the claim. The AC case is analogous.
\end{proof}
We now turn to the case of CS/AC manifolds, concentrating on the situations of
most interest to us.
\begin{decomp}[for CS/AC manifolds]\label{decomp:closedforms_accs}
Let $L$ be a CS/AC manifold with $s$ CS ends and $l$ AC ends. As usual we denote
the union of the CS links by $\Sigma_0$ and the union of the AC links by
$\Sigma_\infty$. Choose a finite-dimensional vector space $H$ of closed 1-forms
on $L$ as in Equation \ref{eq:naturalH}, using $\widetilde{H}_{0,\infty}$ to
denote the space $\widetilde{H}$. For any $i=1,\dots,s+l$ choose a function
$f_i$ such that $f_i\equiv 1$ on the end $S_i$ and $f_i\equiv 0$ on the other
ends. We can assume that $\sum f_i\equiv 1$. Let $E_{0,\infty}$ denote the
$(s+l)$-dimensional vector space generated by these functions. Then, for any
$\boldsymbol{\mu}>0$ and $\boldsymbol{\lambda}<0$,
\begin{equation}\label{eq:closedforms_decay_accs}
\{\mbox{Closed 1-forms on $L$ in
$C^\infty_{(\boldsymbol{\mu}-1,\boldsymbol{\lambda}-1)}(\Lambda^1)$}\}=
\widetilde{H}_{0,\infty}\oplus d(E_{0,\infty})\oplus
d(C^\infty_{(\boldsymbol{\mu},\boldsymbol{\lambda})}(L)).
\end{equation}
Now let $\Lambda^p_{c,\bullet}(L;\R)$ denote the space of $p$-forms on $L$ which
vanish in a neighbourhood of the singularities, with no condition on the large
ends. Let $H^p_{c,\bullet}(L;\R)$ denote the corresponding cohomology groups.
Let $\widetilde{H}^1_{c,\bullet}$ denote the image of the map
$\gamma:H^1_{c,\bullet}(L;\R)\rightarrow H^1(L;\R)$. Choose a finite-dimensional
vector space $\widetilde{H}_{0,\bullet}$ of translation-invariant closed 1-forms
on $L$ with compact support in a neighbourhood of the singularities and such
that the map
\begin{equation}
\widetilde{H}_{0,\bullet}\rightarrow \widetilde{H}^1_{c,\bullet}, \ \
\alpha\mapsto [\alpha]
\end{equation}
is an isomorphism. For any $i=1,\dots,s$ choose a function $f_i$ such that
$f_i\equiv 1$ on the CS end corresponding to the singularity $x_i$ and
$f_i\equiv 0$ on the other ends. Let $E_0$ denote the s-dimensional vector space
generated by these functions. Then, for any $\boldsymbol{\mu}>0$ and
$\boldsymbol{\lambda}>0$,
\begin{equation}\label{eq:closedforms_growth_accs}
\{\mbox{Closed 1-forms on $L$ in
$C^\infty_{(\boldsymbol{\mu}-1,\boldsymbol{\lambda}-1)}(\Lambda^1)$}\}=
\widetilde{H}_{0,\bullet}\oplus d\Big(E_0\oplus
C^\infty_{(\boldsymbol{\mu},\boldsymbol{\lambda})}(L)\Big).
\end{equation}
\end{decomp}
\proof{} The proof is similar to the proofs of the previous decompositions. It
may however be good to emphasize that, in the case $\boldsymbol{\mu}>0$ and
$\boldsymbol{\lambda}>0$, $d(E_0)\cap
d(C^\infty_{(\boldsymbol{\mu},\boldsymbol{\lambda})}(L))\neq\{0\}$ (it is
one-dimensional). This explains the slightly different statement of
Decomposition \ref{eq:closedforms_growth_accs}.
\endproof
\begin{remark}
The weight $\boldsymbol{\beta}=0$ corresponds to an exceptional case in Lemma
\ref{l:formexact}: integration will generally generate log terms, so we cannot
conclude that $A\in C^\infty_{\boldsymbol{\beta}}$ there. One can analogously
argue that
$C^\infty_{-\boldsymbol{1}}(\Lambda^1)/d(C^\infty_{\boldsymbol{0}}(L))$ is not
finite-dimensional.
Similar decompositions hold for $k$-forms: in this setting the exceptional case
corresponds to $\boldsymbol{\beta}=k-1$.
\end{remark}
\begin{remark}
Notice that the above decompositions do not cover all possibilities: for
example, given a CS manifold we could decide to study the space of closed
1-forms in $C^\infty_{\boldsymbol{\beta}-1}(\Lambda^1)$ corresponding to a
weight $\boldsymbol{\beta}=(\beta_1,\dots,\beta_e)$ with some $\beta_i$ positive
and others negative. However, it should be clear from the above discussion how
to use the same ideas to cover any other case of interest. We have restricted
our attention to the cases most relevant to this paper.
\end{remark}
For future reference it is useful to emphasize the topological interpretation of
some of the previous results. The reasons underlying our interest for each case
will become apparent in Section \ref{s:moduli}.
\begin{corollary}\label{cor:topsummary}
Let $L$ be a smooth compact manifold. Then
\begin{equation*}
\{\mbox{Closed $1$-forms on $L$}\}\simeq H^1(L;\R)\oplus d(C^\infty(L)).
\end{equation*}
Let $(L,g)$ be an AC manifold. Then for $\boldsymbol{\beta}<0$
\begin{equation*}
\{\mbox{Closed $1$-forms on $L$ in
$C^\infty_{\boldsymbol{\beta}-1}(\Lambda^1)$}\} \simeq H^1_c(L;\R) \oplus
d(C^\infty_{\boldsymbol{\beta}}(L)),
\end{equation*}
while for $\boldsymbol{\beta}>0$
\begin{equation*}
\{\mbox{Closed $1$-forms on $L$ in
$C^\infty_{\boldsymbol{\beta}-1}(\Lambda^1)$}\} \simeq H^1(L;\R) \oplus
d(C^\infty_{\boldsymbol{\beta}}(L)).
\end{equation*}
Let $(L,g)$ be a CS manifold with link $\Sigma_0$. Then for
$\boldsymbol{\beta}>0$
\begin{align*}
&\{\mbox{Closed $1$-forms on $L$ in
$C^\infty_{\boldsymbol{\beta}-1}(\Lambda^1)$}\}\\
&\quad\simeq \mbox{Ker}\left(H^1(L)\stackrel{\rho}{\rightarrow}
H^1(\Sigma_0)\right) \oplus d(E_0)\oplus d(C^\infty_{\boldsymbol{\beta}}(L)).
\end{align*}
Let $(L,g)$ be a CS/AC manifold with link $\Sigma=\Sigma_0\amalg\Sigma_\infty$.
Then for $\boldsymbol{\mu}>0$ and $\boldsymbol{\lambda}<0$
\begin{align*}
&\{\mbox{Closed $1$-forms on $L$ in
$C^\infty_{(\boldsymbol{\mu}-1,\boldsymbol{\lambda}-1)}(\Lambda^1)$}\}\\
&\quad\simeq \mbox{Ker}\left(H^1_{\bullet,c}(L)\stackrel{\rho}{\rightarrow}
H^1(\Sigma_0)\right)\oplus d(E_0) \oplus
d(C^\infty_{(\boldsymbol{\mu},\boldsymbol{\lambda})}(L)),
\end{align*}
while for $\boldsymbol{\mu}>0$ and $\boldsymbol{\lambda}>0$
\begin{align*}
&\{\mbox{Closed $1$-forms on $L$ in
$C^\infty_{(\boldsymbol{\mu}-1,\boldsymbol{\lambda}-1)}(\Lambda^1)$}\}\\
&\quad\simeq \mbox{Ker}\left(H^1(L)\stackrel{\rho}{\rightarrow}
H^1(\Sigma_0)\right)\oplus d\left(E_0\oplus
C^\infty_{(\boldsymbol{\mu},\boldsymbol{\lambda})}(L)\right).
\end{align*}
\end{corollary}
\begin{proof}
The compact case coincides with Equation \ref{eq:closedforms}. The AC case with
$\boldsymbol{\beta}<0$ follows from Equation \ref{eq:closedforms_decay_ac} and
Remark \ref{rem:compactsupport}. The AC case with $\boldsymbol{\beta}>0$
coincides with Equation \ref{eq:closedforms_growth_ac}. The CS case coincides
with Equation \ref{eq:closedforms_decay_cs}.
Let us now focus on the CS/AC case with $\boldsymbol{\lambda}<0$. Using the
notation of Decomposition \ref{decomp:closedforms_accs}, let $E'$ denote a
complement of $E_0\oplus\R$ in $E_{0,\infty}$, \textit{i.e.}
$E_{0,\infty}=E_0\oplus\R\oplus E'$. Notice that the long exact sequence
\ref{eq:cohomsequence} with $\Sigma=\Sigma_0\amalg\Sigma_\infty$ leads to an
identification $H^1_c(L;\R)\simeq \widetilde{H}^1_c(L)\oplus d(E_{0,\infty})$.
One can also set up the ``relative" analogue of Sequence \ref{eq:cohomsequence}
using the inclusion of pairs $(\Sigma_0,\emptyset)\subset (L,\Sigma_\infty)$.
Using notation analogous to that of Decomposition \ref{decomp:closedforms_accs}
this leads to the long exact sequence
\begin{equation*}
0\rightarrow H^0_c(L;\R)\rightarrow H^0_{\bullet,c}(L;\R)\rightarrow
H^0(\Sigma_0;\R)\rightarrow H^1_c(L;\R)\stackrel{\gamma}{\rightarrow}
H^1_{\bullet,c}(L;\R)\stackrel{\rho}{\rightarrow}
H^1(\Sigma_0;\R)\rightarrow\dots
\end{equation*}
Since $H^0_c(L;\R)=0$ and $H^0_{\bullet,c}(L;\R)=0$, one obtains an
identification $H^1_c(L;\R)\simeq
E_0\oplus\mbox{Ker}\left(H^1_{\bullet,c}(L)\stackrel{\rho}{\rightarrow}
H^1(\Sigma_0)\right)$. Comparing these identifications yields an identification
$\widetilde{H}^1_c(L;\R)\oplus
d(E')\simeq\mbox{Ker}\left(H^1_{\bullet,c}(L)\stackrel{\rho}{\rightarrow}
H^1(\Sigma_0)\right)$. The claim follows.
Now consider the CS/AC case with $\boldsymbol{\lambda}>0$. The long exact
sequence \ref{eq:cohomsequence} with $\Sigma=\Sigma_0$ yields
\begin{equation}\label{eq:cohomsequence_accsbis}
0\rightarrow H^0(L;\R)\rightarrow H^0(\Sigma_0;\R)\rightarrow
H^1_{c,\bullet}(L;\R)\stackrel{\gamma}{\rightarrow}
H^1(L;\R)\stackrel{\rho}{\rightarrow} H^1(\Sigma_0;\R)\rightarrow\dots
\end{equation}
This proves the final claim.
\end{proof}
\section{Lagrangian conifolds}\label{s:lagconifolds}
A priori, a \textit{CS/AC submanifold} might simply be defined as an immersed
submanifold whose topology and induced metric is of the type defined in Section
\ref{ss:accs_review}. However, for the purposes of this article it is convenient
to strengthen the hypotheses by adding the requirement that the submanifold have
a well-defined cone at each singularity and at each end. The precise definitions
are as follows. We restrict our attention to Lagrangian submanifolds in K\"ahler
ambient spaces, but it is clear how one might extend these definitions to other
settings.
\begin{definition} \label{def:lagr}
Let $(M^{2m},\omega)$ be a symplectic manifold. An embedded or immersed
submanifold $\iota:L^m\rightarrow M$ is \textit{Lagrangian} if
$\iota^*\omega\equiv 0$. The immersion allows us to view the tangent bundle $TL$
of $L$ as a subbundle of $TM$ (more precisely, of $\iota^*TM$). When $M$ is
K\"ahler with structures $(g,J,\omega)$ it is simple to check that $L$ is
Lagrangian iff $J$ maps $TL$ to the normal bundle $NL$ of $L$, \textit{i.e.}
$J(TL)=NL$.
\end{definition}
\begin{definition}\label{def:aclagsub}
Let $L^m$ be a smooth manifold. Assume given a Lagrangian immersion
$\iota:L\rightarrow \C^m$, the latter endowed with its standard structures
$\tilde{J},\tilde{\omega}$. We say that $L$ is an \textit{asymptotically conical Lagrangian submanifold}
with \textit{rate} $\boldsymbol{\lambda}$ if it satisfies the following
conditions.
\begin{enumerate}
\item We are given a compact subset $K\subset L$ such that $S:=L\setminus K$ has
a finite number of connected components $S_1,\dots,S_e$.
\item We are given Lagrangian cones $\mathcal{C}_i\subset \C^m$ with smooth
connected links $\Sigma_i:=\mathcal{C}_i\bigcap \Sph^{2m-1}$. Let
$\iota_i:\Sigma_i\times (0,\infty)\rightarrow \C^m$ denote the natural
immersions, parametrizing $\mathcal{C}_i$.
\item We are finally given an $e$-tuple of \textit{convergence rates}
$\boldsymbol{\lambda}=(\lambda_1,\dots,\lambda_e)$ with $\lambda_i<2$, \textit{centers} $p_i\in\C^m$ and
diffeomorphisms $\phi_i:\Sigma_i\times [R,\infty)\rightarrow \overline{S_i}$
for some $R>0$ such that, for $r\rightarrow\infty$ and all $k\geq 0$,
\begin{equation}\label{eq:aclagdecay}
|\tnabla^k(\iota\circ\phi_i-(\iota_i+p_i)|=O(r^{\lambda_i-1-k})
\end{equation}
with respect to the conical metric $\tg_i$ on $\cone_i$.
\end{enumerate}
\end{definition}
Notice that the restriction $\lambda_i<2$ ensures that the cone is unique but is
weak enough to allow the submanifold to converge to a translated copy
$\mathcal{C}_i+p'_i$ of the cone (\textit{e.g.} if $\lambda_i=1$), or even to
slowly pull away from the cone (if $\lambda_i>1$).
\begin{definition}\label{def:cslagsub}
Let $\bar{L}^m$ be a smooth manifold except for a finite number of possibly singular
points $\{x_1,\dots,x_e\}$. Assume given a continuous map
$\iota:\bar{L}\rightarrow \C^m$ which restricts to a smooth Lagrangian immersion of
$L:=\bar{L}\setminus\{x_1,\dots,x_e\}$. We say that $\bar{L}$ (or $L$) is a \textit{conically singular
Lagrangian submanifold} with \textit{rate} $\boldsymbol{\mu}$ if it satisfies
the following conditions.
\begin{enumerate}
\item We are given open connected neighbourhoods $S_i$ of $x_i$.
\item We are given Lagrangian cones $\mathcal{C}_i\subset \C^m$ with smooth
connected links $\Sigma_i:=\mathcal{C}_i\bigcap \Sph^{2m-1}$. Let
$\iota_i:\Sigma_i\times (0,\infty)\rightarrow \C^m$ denote the natural
immersions, parametrizing $\mathcal{C}_i$.
\item We are finally given an $e$-tuple of \textit{convergence rates}
$\boldsymbol{\mu}=(\mu_1,\dots,\mu_e)$ with $\mu_i>2$, \textit{centers} $p_i\in\C^m$ and diffeomorphisms
$\phi_i:\Sigma_i\times (0,\epsilon]\rightarrow \overline{S_i}\setminus\{x_i\}$ such that, for
$r\rightarrow 0$ and all $k\geq 0$,
\begin{equation}\label{eq:cslagdecay}
|\tnabla^k(\iota\circ\phi_i-(\iota_i+p_i))|=O(r^{\mu_i-1-k})
\end{equation}
with respect to the conical metric $\tg_i$ on $\cone_i$. Notice that our assumptions imply that $\iota(x_i)=p_i$.
\end{enumerate}
\end{definition}
It is simple to check that AC Lagrangian submanifolds, with the induced metric,
satisfy Definition \ref{def:ac_manifold} with $\nu_i=\lambda_i-2$. The analogous
fact holds for CS Lagrangian submanifolds.
\begin{definition} \label{def:accslagsub}
Let $\bar{L}^m$ be a smooth manifold except for a finite number of possibly singular
points $\{x_1,\dots,x_s\}$ and with $l$ ends. Assume
given a continuous map $\iota:\bar{L}\rightarrow \C^m$ which restricts to a smooth Lagrangian
immersion of $L:=\bar{L}\setminus\{x_1,\dots,x_s\}$. We say that $\bar{L}$ (or
$L$) is a \textit{CS/AC Lagrangian submanifold} with \textit{rate}
$(\boldsymbol{\mu},\boldsymbol{\lambda})$ if in a neighbourhood of the points
$x_i$ it has the structure of a CS submanifold with rates $\mu_i$ and in a
neighbourhood of the remaining ends it has the structure of an AC submanifold
with rates $\lambda_i$.
We use the generic term \textit{Lagrangian conifold} (even though ``subconifold"
would be more appropriate) to indicate any CS, AC or CS/AC Lagrangian
submanifold.
\end{definition}
\begin{example} \label{e:lagcone}
Let $\mathcal{C}$ be a cone in $\C^m$ with smooth link $\Sigma^{m-1}$. It can be
shown that $\mathcal{C}$ is a Lagrangian iff $\Sigma$ is \textit{Legendrian} in
$\Sph^{2m-1}$ with respect to the natural \textit{contact structure} on the
sphere. Then $\mathcal{C}$ is a CS/AC Lagrangian submanifold of $\C^m$ with rate
$(\boldsymbol{\mu,\lambda})$ for any $\boldsymbol{\mu}$ and
$\boldsymbol{\lambda}$.
\end{example}
The definition of CS Lagrangian submanifolds can be generalized to K\"ahler
ambient spaces as follows. Once again we denote the standard structures on
$\C^m$ by $\tilde{J}$, $\tilde{\omega}$.
\begin{definition} \label{def:generalcslagsub}
Let $(M^{2m},J,\omega)$ be a K\"ahler manifold and $\bar{L}^m$ be a smooth
manifold except for a finite number of possibly singular points $\{x_1,\dots,x_e\}$. Assume given a continuous map $\iota:\bar{L}\rightarrow M$ which restricts
to a smooth Lagrangian immersion of $L:=\bar{L}\setminus\{x_1,\dots,x_e\}$. We say that $\bar{L}$ (or
$L$) is a \textit{Lagrangian submanifold with conical singularities} (CS
Lagrangian submanifold) if it satisfies the following conditions.
\begin{enumerate}
\item We are given isomorphisms $\upsilon_i:\C^m\rightarrow T_{\iota(x_i)}M$ such that
$\upsilon_i^*\omega=\tilde{\omega}$ and $\upsilon_i^*J=\tilde{J}$.
According to Darboux' theorem, cf. \textit{e.g.} \cite{weinstein}, there then
exist an open ball $B_R$ in $\C^m$ (of small radius $R$) and diffeomorphisms
$\Upsilon_i:B_R\rightarrow M$ such that $\Upsilon(0)=\iota(x_i)$,
$d\Upsilon_i(0)=\upsilon_i$ and $\Upsilon_i^*\omega=\tilde{\omega}$.
\item We are given open neighbourhoods $S_i$ of $x_i$ in $\bar{L}$. We assume $S_i$ are small, in the sense that
the compositions
\begin{equation*}
\Upsilon_i^{-1}\circ\iota:S_i\rightarrow B_R
\end{equation*}
are well-defined.
We are
also given Lagrangian cones $\mathcal{C}_i\subset\C^m$ with smooth connected
links $\Sigma_i:=\mathcal{C}_i\bigcap \Sph^{2m-1}$. Let $\iota_i:\Sigma_i\times
(0,\infty)\rightarrow \C^m$ denote the natural immersions, parametrizing
$\cone_i$.
\item We are finally given an $e$-tuple of \textit{convergence rates}
$\boldsymbol{\mu}=(\mu_1,\dots,\mu_e)$ with $\mu_i\in (2,3)$ and diffeomorphisms
$\phi_i:\Sigma_i\times (0,\epsilon]\rightarrow \overline{S_i}\setminus\{x_i\}$ such that, as $r\rightarrow 0$ and
for all $k\geq 0$,
\begin{equation}\label{eq:generalcslagdecay}
|\tnabla^k(\Upsilon_i^{-1}\circ\iota\circ\phi_i-\iota_i)|=O(r^{\mu_i-1-k})
\end{equation}
with respect to the conical metric $\tg_i$ on $\cone_i$.
\end{enumerate}
We call $x_i$ the \textit{singularities} of $\bar{L}$ and $\upsilon_i$ the
\textit{identifications}.
\end{definition}
One can check that, when $M=\C^m$, Definition \ref{def:generalcslagsub} coincides with Definition \ref{def:cslagsub} if we choose $\Upsilon_i(x):=x+\iota(x_i)$. Notice that the local identifications of $M$ with $\C^m$ are prescribed only up to first order: we correspondingly introduce a constraint on the range of $\mu_i$ to ensure that the rate depends only on $\upsilon_i$, not on $\Upsilon_i$.
\begin{remark}
One could also define and study AC Lagrangian submanifolds in $M$, but this
would require a preliminary study of AC metrics on K\"ahler manifolds, going
beyond the scope of this article. We refer to \cite{pacini:defs} for some
details in this direction.
\end{remark}
\section{Deformations of Lagrangian conifolds}\label{s:lagdefs}
We now want to understand how to parametrize the \textit{Lagrangian
deformations} of a given Lagrangian conifold $L\subset M$. Since the Lagrangian
condition is invariant under reparametrization of $L$, to avoid huge amounts of
geometric redundancy it is best to work in terms of non-parametrized
submanifolds; in other words, in terms of equivalence classes of immersed
submanifolds, where two immersions are \textit{equivalent} if they differ by a
reparametrization. Then, to parametrize the possible deformations of $L$, it is
sufficient to prove a \textit{Lagrangian neighbourhood theorem}.
\begin{remark}
The analogous situation in the Riemannian setting is well-known. The set
$\mbox{Imm}(L,M)$ of immersions $L\rightarrow (M,g)$ can be topologized via the
\textit{$C^1$} or \textit{Whitney} topology, \textit{i.e.} in terms of the
natural topology on the first jet bundle $J^1(L,M)$. The group of
diffeomorphisms $\mbox{Diff}(L)$ acts on this space by reparametrization. Choose
an element $\iota\in \mbox{Imm}(L,M)$. Let $NL$ denote the normal bundle. Using
the \textit{tubular neighbourhood theorem} one can define a natural injection
$$\Lambda^0(NL)\rightarrow\mbox{Imm}(L,M)/\mbox{Diff}(L).$$
In standard situations (for example when $L$ is compact) this actually defines a
local homeomorphism between the natural topologies on these spaces.
\end{remark}
A foundation for the theory of Lagrangian neighbourhoods is provided by the
following linear-algebraic construction. Let $W$ be a finite-dimensional real
vector space. Then $W\oplus W^*$ admits a canonical symplectic structure
$\hat{\omega}$ defined as follows:
\begin{equation}\label{eq:canonicallinearcotg}
\hat{\omega}(w_1+\alpha_1,w_2+\alpha_2):=\alpha_2(w_1)-\alpha_1(w_2).
\end{equation}
It turns out that this example of symplectic vector space is actually very
general, in the following sense. Let $(V,\omega)$ be a symplectic vector space.
Let $W\subset V$ be a Lagrangian subspace. Choose a Lagrangian complement
$Z\subset V$, so that $V=W\oplus Z$. It is simple to check that the restriction
of $\omega$ to $Z$ defines an isomorphism
\begin{equation}\label{eq:canonicalisgeneral}
\omega_{|Z}:Z\rightarrow W^*,\ \ z\mapsto\omega(z,\cdot)
\end{equation}
and that, using this isomorphism, one can build an isomorphism $\gamma:(W\oplus
W^*,\hat{\omega})\simeq (V,\omega)$. Furthermore, such $\gamma$ is unique if we
impose that it coincide with the identity on $W$. Adding this condition thus
implies that $\gamma$ is uniquely defined by the choice of $Z$.
It is a well-known fact, first noticed by Souriau \cite{souriau}, that a similar
construction exists also for symplectic manifolds. The construction is based on
the following standard facts. Given any manifold $L$, the cotangent bundle
$T^*L$ admits a canonical symplectic structure $\hat{\omega}$. Specifically,
consider the \textit{tautological} 1-form on $T^*L$ defined by
$\hat{\lambda}[\alpha](v):=\alpha(\pi_*(v))$, where $\pi:T^*L\rightarrow L$ is
the natural projection. Then $\hat{\omega}:=-d\hat{\lambda}$. Notice that a
section of $T^*L$ is simply a 1-form $\alpha$ on $L$. The graph $\Gamma(\alpha)$
is Lagrangian in $T^*L$ iff $\alpha$ is closed. In particular the zero section
$L\subset T^*L$ is Lagrangian. Furthermore each fibre $\pi^{-1}(p)=T^*_pL$ is a
Lagrangian submanifold. The fibres thus define a Lagrangian foliation of $T^*L$
transverse to the zero section. Finally, every 1-form $\alpha$ defines a
\textit{translation} map
\begin{equation}\label{eq:translation}
\tau_\alpha:T^*L\rightarrow T^*L,\ \ \tau_\alpha(x,\eta):=(x,\alpha(x)+\eta).
\end{equation}
If $\alpha$ is closed then this map is a symplectomorphism of
$(T^*L,\hat{\omega})$.
\subsection{First case: smooth compact Lagrangian
submanifolds}\label{ss:cptlagdefs}
We can now quote Souriau's result, following Weinstein \cite{weinstein}
Corollary 6.2.
\begin{theorem}\label{th:nbd_weinstein}
Let $(M,\omega)$ be a symplectic manifold. Let $L\subset M$ be a smooth compact
Lagrangian submanifold. Then there exist a neighbourhood $\mathcal{U}$ of the
zero section of $L$ inside its cotangent bundle $T^*L$ and an embedding
$\Phi_L:\mathcal{U}\rightarrow M$ such that $\Phi_{L|L}=Id:L\rightarrow L$ and
$\Phi_L^*\omega=\hat{\omega}$.
\end{theorem}
\begin{proof}
For each $x\in L$, $T_xL$ is a Lagrangian subspace of $T_xM$. The first step is
to choose a Lagrangian complement $Z_x$, so that $T_xM=T_xL\oplus Z_x$. This can
be done smoothly with respect to $x$ using the fact that the space of Lagrangian
complements is a contractible set inside the Grassmannian of $m$-planes in
$T_xM$. As seen following Equation \ref{eq:canonicalisgeneral}, $\omega$ then
provides an isomorphism $\gamma_x:(T_xL\oplus T_x^*L,\hat{\omega})\rightarrow
(T_xM,\omega)$, uniquely defined by the condition that $\gamma_x=Id$ on $T_xL$.
Now choose a diffeomorphism $\Psi_L:\mathcal{U}\rightarrow M$ such that
$(\Psi_L)_*$ extends $\gamma$. By construction, the pull-back form
$(\Psi_L)^*\omega$ coincides with $\hat{\omega}$ at each point of $L$. We now
need to perturb $\Psi_L$ so that the pull-back form coincides with
$\hat{\omega}$ in a neighbourhood of $L$. Set $\omega_0:=\hat{\omega}$ and
$\omega_1:=(\Psi_L)^*\omega$. One can use an argument due to Moser together with
the Poincar\'e Lemma to prove that there exists a diffeomorphism
$k:T^*L\rightarrow T^*L$ such that $k^*\omega_1=\omega_0$ and
$k_{|L}=Id:L\rightarrow L$. Thus $\Phi_L:=\Psi_L\circ k$ has the required
properties. For later use it is also useful to note that, using the same
argument as in \cite{weinstein} Theorem 7.1, one can further show that, at each
$x\in L$, $k_*$ preserves $T_x^*L$. A linear-algebraic argument then shows that
this implies that $k_*=Id$ at each $x\in L$. Thus $(\Phi_L)_*=(\Psi_L)_*$ at
each $x\in L$.
\end{proof}
\begin{remark} Although the statement and proof are for embedded submanifolds it
is not difficult to extend them to immersed compact Lagrangian submanifolds by
working locally. In this case $\Phi_L$ will only be a local embedding.
\end{remark}
Let $C^\infty(\mathcal{U})$ denote the space of smooth 1-forms on $L$ whose
graph lies in $\mathcal{U}$. Theorem \ref{th:nbd_weinstein} leads immediately to
the following conclusion.
\begin{corollary} \label{cor:nbd_weinstein}
Let $(M,\omega)$ be a symplectic manifold. Let $L\subset M$ be a smooth compact
Lagrangian submanifold. Then $\Phi_L$ defines by composition an injective map
\begin{equation}\label{eq:cptlagrcorrespondence}
\Phi_L:C^\infty(\mathcal{U})\rightarrow \mbox{Imm}(L,M)/\mbox{Diff}(L).
\end{equation}
A section $\alpha\in C^\infty(\mathcal{U})$ is closed iff the corresponding
(non-parametrized) submanifold $\Phi_L\circ\alpha$ is Lagrangian.
\end{corollary}
An important point about the map $\Phi_L$ in Equation
\ref{eq:cptlagrcorrespondence} is that any submanifold which admits a
parametrization which is $C^1$-close to some parametrization of $L$ belongs to
the image of $\Phi_L$, \textit{i.e.} corresponds to a 1-form $\alpha$.
Let $\mbox{Lag}(L,M)$ denote the set of Lagrangian immersions from $L$ into $M$.
Using Corollary \ref{cor:nbd_weinstein} and the Fr\'echet topology on
$C^\infty(\mathcal{U})$ we can locally define a topology on
$\mbox{Lag}(L,M)/\mbox{Diff}(L)$; one can then check that on the intersection of
any two open sets these topologies coincide, so we obtain a global topology on
$\mbox{Lag}(L,M)/\mbox{Diff}(L)$. The connected component containing the given
$L\subset M$ defines the \textit{moduli space of Lagrangian deformations of
$L$}. Coupling Corollary \ref{cor:nbd_weinstein} with Decomposition
\ref{decomp:closedforms} gives a good idea of the local structure of this space.
\subsection{Second case: Lagrangian cones in $\C^m$}\label{ss:conelagdefs}
Let $\mathcal{C}$ be a Lagrangian cone in $\C^m$ with link $(\Sigma,g')$ and
conical metric $\tg$. The goal of this section is to provide an analogue of the
theory of Section \ref{ss:cptlagdefs} for this specific submanifold, giving a correspondence between closed 1-forms in $C^\infty_{(\mu-1,\lambda-1)}(\Lambda^1)$ and Lagrangian deformations of $\mathcal{C}$ with rate $(\mu,\lambda)$.
Let $\theta$ denote the generic point on $\Sigma$. We will identify
$\Sigma\times (0,\infty)$ with $\mathcal{C}$ via the immersion
\begin{equation}
\iota:\Sigma\times (0,\infty)\rightarrow \C^m, \ \ (\theta,r)\mapsto r\theta.
\end{equation}
\begin{remark}\label{rem:iota_*}
Let $\theta(t)$ be a curve in $\Sigma$ such that $\theta(0)=\theta$. Let $r(t)$
be a curve in $\R^+$ such that $r(0)=r$.
Differentiating $\iota$ at the point $(\theta,r)$ gives identifications
\begin{equation}
\begin{array}{rcl}
\iota_*:T_\theta\Sigma\oplus\R & \rightarrow &
T_{r\theta}\mathcal{C}\subset\C^m\\
(\theta'(0),r'(0)) & \mapsto &
d/dt\left(r(t)\theta(t)\right)_{|t=0}=r'(0)\theta+r\theta'(0)\in\C^m.
\end{array}
\end{equation}
This leads to the general formula $\iota_{*|\theta,r}(v,a)=a\theta+rv$.
\end{remark}
We can build an explicit (local) identification $\Psi_{\mathcal{C}}$ of
$T^*\mathcal{C}$ with $\C^m$ as follows.
Firstly, the metric $\tg$ gives an identification
\begin{equation}\label{eq:cotg=tg}
T^*\mathcal{C}\rightarrow T\mathcal{C},\ \
(\theta,r,\alpha_1+\alpha_2\,dr)\mapsto (\theta,r,r^{-2}A_1+\alpha_2\partial r),
\end{equation}
where $g'(A_1,\cdot)=\alpha_1$ and we use the notation of Section
\ref{ss:closedforms}. Notice that, according to Remark \ref{rem:iota_*}, the
corresponding vector in $\C^m$ is $\iota_*(r^{-2}A_1+\alpha_2\partial
r)=\alpha_2\theta+r^{-1}A_1$. Notice also that Equation \ref{eq:cotg=tg} defines
a fibrewise isometry between vector bundles over $\mathcal{C}$. Let
$\widetilde{\nabla}$ denote the standard connection on the tangent bundle of
$\C^m$. Since $\mathcal{C}$ has the induced metric, the Levi-Civita connection
on $T\mathcal{C}$ coincides with the tangential projection
$\widetilde{\nabla}^T$. Let $T^*\mathcal{C}$ have the induced Levi-Civita
connection. Then Equation \ref{eq:cotg=tg} also defines an isomorphism between
the two connections.
Secondly, since $\mathcal{C}$ is Lagrangian the complex structure provides an
identification
\begin{equation}\label{eq:tg=normal}
\tilde{J}:T\mathcal{C}\simeq N\mathcal{C}.
\end{equation}
This is again a fibrewise isometry. The perpendicular component
$\widetilde{\nabla}^\perp$ defines a connection on $N\mathcal{C}$. Since $\C^m$
is K\"ahler, $\widetilde{\nabla} \tilde{J}=\tilde{J}\widetilde{\nabla}$. Thus
$\widetilde{\nabla}^\perp \tilde{J}=\tilde{J}\widetilde{\nabla}^T$, so Equation
\ref{eq:tg=normal} defines an isomorphism between the two connections.
Thirdly, the Riemannian tubular neighbourhood theorem gives an explicit (local)
identification
\begin{equation}\label{eq:normal=ambient}
N\mathcal{C}\rightarrow \C^m,\ \ v\in N_{r\theta}\mathcal{C}\mapsto r\theta+v.
\end{equation}
By composition we now obtain the required identification
\begin{equation}\label{eq:psicone}
\Psi_{\mathcal{C}}:\mathcal{U}\subset T^*\mathcal{C}\rightarrow\C^m,\ \
(\theta,r,\alpha_1+\alpha_2\,dr)\mapsto
r\theta+\tilde{J}(\alpha_2\theta+r^{-1}A_1).
\end{equation}
Now let $\alpha$ be a 1-form on $\mathcal{C}$. Then, under the above
identifications, $(\Psi_{\mathcal{C}}\circ\alpha)-\iota\simeq\alpha$. This shows
that if $\alpha\in C^\infty_{(\mu-1,\lambda-1)}(\mathcal{U})$ for some $\mu>2$,
$\lambda<2$ then $\Psi_{\mathcal{C}}\circ\alpha$ is a CS/AC submanifold in
$\C^m$ asymptotic to $\mathcal{C}$ with rate $(\mu,\lambda)$.
Notice also that
\begin{equation}\label{eq:psiequivariant}
\Psi_{\mathcal{C}}(\theta,tr,t^2\alpha_1+t\alpha_2\,dr)=t\Psi_{\mathcal{C}}
(\theta,r,\alpha_1+\alpha_2\,dr).
\end{equation}
This suggests that we define an action of $\R^+$ on $T^*\mathcal{C}$ as follows:
\begin{equation}\label{eq:action}
\R^+\times T^*\mathcal{C}\rightarrow T^*\mathcal{C},\ \
t\cdot(\theta,r,\alpha_1+\alpha_2\,dr):=(\theta,tr,t^2\alpha_1+t\alpha_2\,dr).
\end{equation}
With respect to this action on $T^*\mathcal{C}$ and the standard action by
dilations on $\C^m$, Equation \ref{eq:psiequivariant} shows that
$\Psi_{\mathcal{C}}$ is an equivariant map.
\begin{remark} \label{rem:action}
Equation \ref{eq:action} introduces an action on $T^*\mathcal{C}$ which rescales
both the base space and the fibres. We can also obtain it as follows. On any
cotangent bundle $T^*L$ there is a natural action
\begin{equation*}
\R^+\times T^*L\rightarrow T^*L,\ \ t\cdot(x,\alpha):=(x,t^2\alpha).
\end{equation*}
The induced action on 1-forms is such that, for the tautological 1-form
$\hat{\lambda}$, $t^*\hat{\lambda}=t^2\hat{\lambda}$.
When $L=\Sigma\times (0,\infty)$ there is also a natural action
\begin{equation*}
\R^+\times L\rightarrow L,\ \ t\cdot(\theta,r):=(\theta,tr).
\end{equation*}
This induces an action on $T^*L$ as follows:
\begin{equation*}
\R^+\times T^*(\Sigma\times (0,\infty))\rightarrow
T^*(\Sigma\times(0,\infty)),\ \
t\cdot(\theta,r,\alpha_1+\alpha_2\,dr):=(\theta,tr,\alpha_1+t^{-1}\alpha_2\,dr).
\end{equation*}
The induced action on 1-forms preserves
$\hat{\lambda}:t^*\hat{\lambda}=\hat{\lambda}$. Equation \ref{eq:action}
coincides with the composed action and thus satisfies
$t^*\hat{\lambda}=t^2\hat{\lambda}$, so $t^*\hat{\omega}=t^2\hat{\omega}$.
\end{remark}
We now want to investigate the symplectic properties of the map
$\Psi_{\mathcal{C}}$. Let $\tilde{\omega}$ denote the standard symplectic
structure on $\C^m$. Since $\mathcal{C}$ is Lagrangian, the fibres of the normal
bundle define (locally) a Lagrangian foliation of $\C^m$, transverse to
$\mathcal{C}$. Using the fact that $\Psi_{\mathcal{C}}$ is the identity on
$\mathcal{C}$ and is linear on each fibre, one can check that, at each point of
$\mathcal{C}$, $(\Psi_{\mathcal{C}})^*\tilde{\omega}=\hat{\omega}$. Notice also
that $\Psi_{\mathcal{C}}$ identifies the foliation of $\C^m$ with the foliation
of $T^*\mathcal{C}$ defined by the fibres.
As in the proof of Theorem \ref{th:nbd_weinstein}, we now want to perturb
$\Psi_{\mathcal{C}}$ so as to obtain a local symplectomorphism
$\mathcal{U}\subset T^*\mathcal{C}\rightarrow \C^m$. As in that case, the idea
is to build a (local) diffeomorphism $k:T^*\mathcal{C}\rightarrow
T^*\mathcal{C}$ such that $k_*=Id$ at each point of $\mathcal{C}$ and
$k^*(\Psi_{\mathcal{C}})^*\tilde{\omega}=\hat{\omega}$. The construction of such
$k$ is sufficiently explicit in \cite{weinstein} p. 333 to allow us to prove
that $k$ is equivariant with respect to the $\R^+$-action. Furthermore, the fact
that the fibres of $T^*\mathcal{C}$ are Lagrangian for both symplectic forms
implies that $k$ preserves these fibres, see \cite{weinstein} Theorem 7.1 for
details. Now define
\begin{equation}\label{eq:lagconemap}
\Phi_{\mathcal{C}}:=\Psi_{\mathcal{C}}\circ k:\mathcal{U}\subset
T^*\mathcal{C}\rightarrow \C^m.
\end{equation}
By construction, $\Phi_{\mathcal{C}}$ satisfies
$(\Phi_{\mathcal{C}})^*\tilde{\omega}=\hat{\omega}$. Furthermore,
$\Phi_{\mathcal{C}}$ is equivariant and its fibrewise linearization at each
$x\in\mathcal{C}$ coincides with $\Psi_{\mathcal{C}}$. Thus
$\Phi_{\mathcal{C}}=\Psi_{\mathcal{C}}+R$, for some $R$ satisfying
\begin{equation}
|R(\theta,1,\alpha_1,\alpha_2)|=O(|\alpha_1|_{g'}^2+|\alpha_2|^2),\ \ \mbox{ as }|a_1|_{g'}+|a_2|\rightarrow 0.
\end{equation}
Clearly $R$ is also equivariant. Thus
\begin{equation}\label{eq:Requivariance}
|R(\theta,t,\alpha_1,\alpha_2)|=|R(\theta,t\cdot
1,t^2t^{-2}\alpha_1,tt^{-1}\alpha_2)|=t\cdot
O(t^{-4}|\alpha_1|_{g'}^2+t^{-2}|\alpha_2|^2).
\end{equation}
The equivariance of R can be used to determine its asymptotic
behaviour with respect to $r$ after composition with 1-forms on $\mathcal{C}$. For example, given
any $\mu>2$ and $\lambda<2$, choose $\alpha$ in the space
$C^\infty_{(\mu-1,\lambda-1)}(\mathcal{U})$. Notice that, as
$r\rightarrow\infty$, $r^{-1}|\alpha_1|_{g'}=|\alpha_1|_g=O(r^{\lambda-1})$.
This implies $r^{-4}|\alpha_1|_{g'}^2=O(r^{2\lambda-4})$. Analogously,
$|\alpha_2|=O(r^{\lambda-1})$ so $r^{-2}|\alpha_2|^2=O(r^{2\lambda-4})$.
Equation \ref{eq:Requivariance} then shows that
$(R\circ\alpha)(\theta,r)=R(\theta,r,\alpha(\theta,r))$ satisfies $|R\circ\alpha|=O(r^{2\lambda-3})$ as $r\rightarrow \infty$. Further calculations show that
the derivatives of $R\circ\alpha$ scale correspondingly, \textit{e.g.}
\begin{equation}\label{eq:Rscaling}
|(R\circ\alpha)_*(\partial r)|=O(r^{2\lambda-4}),\ \
|(R\circ\alpha)_*(r^{-1}\partial\theta_i)|=O(r^{2\lambda-4}).
\end{equation}
More generally, $|\tnabla^k(R\circ\alpha)|=O(r^{2\lambda-3-k})$. As a result,
\begin{equation*}
\begin{split}
|\tnabla^k(\Phi_{\mathcal{C}}\circ\alpha-\iota)| &= |\tnabla^k((\Psi_{\mathcal{C}}+R)\circ\alpha-\iota)|
\leq |\tnabla^k(\Psi_{\mathcal{C}}\circ\alpha-\iota)|+|\tnabla^k(R\circ\alpha)|\\
&= O(r^{\lambda-1-k})+O(r^{2\lambda-3-k})=O(r^{\lambda-1-k}),
\end{split}
\end{equation*}
where we use $\lambda<2$. This shows that $\Phi_{\mathcal{C}}\circ \alpha$ is a CS/AC
Lagrangian submanifold asymptotic to $\mathcal{C}$ with rate $(\mu,\lambda)$.
Conversely, one can show that any Lagrangian submanifold $L$ of $\C^m$ which
admits a parametrization which is $C^1$-close to $\iota$ and which is asymptotic
to $\iota$ in the sense of Equations \ref{eq:aclagdecay} and \ref{eq:cslagdecay}
corresponds to a closed 1-form $\alpha\in
C^\infty_{(\mu-1,\lambda-1)}(\mathcal{U})$.
In complete analogy with Section \ref{ss:cptlagdefs} we can use
$\Phi_{\mathcal{C}}$ and the closed forms in the space
$C^\infty_{(\mu-1,\lambda-1)}(\mathcal{U})$ to define a topology on the set of
Lagrangian submanifolds which admit a parametrization
$\iota:\Sigma\times\R^+\rightarrow\C^m$ which is asymptotic to $\mathcal{C}$
with rate $(\mu,\lambda)$. The connected component containing $\mathcal{C}$
defines the \textit{moduli space of CS/AC Lagrangian deformations of
$\mathcal{C}$ with rate $(\mu,\lambda)$}.
We conclude with a last comment on the differential properties of
$\Psi_{\mathcal{C}}$. Recall the following general fact.
\begin{lemma} \label{l:diffvsconn}
Let $E\rightarrow M$ be a vector bundle, endowed with a connection $\nabla$. Let
$\sigma:M\rightarrow E$ be a section of $E$. Choose $v\in T_pM$. The connection
defines a decomposition into ``vertical" and ``horizontal" subspaces
\begin{equation}
T_{\sigma(p)}E=V_{\sigma(p)}\oplus H_{\sigma(p)}, \mbox{\ \ with\ \
}V_{\sigma(p)}\simeq E_p,\ H_{\sigma(p)}\simeq T_p(M).
\end{equation}
Under these identifications, $\sigma_*(v)\simeq \nabla_v\sigma+v$.
\end{lemma}
We can apply Lemma \ref{l:diffvsconn} as follows. Let $\alpha$ be a section of
$T^*\mathcal{C}$ so that $\Psi_{\mathcal{C}}\circ\alpha:\mathcal{C}\rightarrow
\C^m$ is a submanifold of $\C^m$. Choose $v\in T_{r\theta}\mathcal{C}$. Then,
using the identifications \ref{eq:cotg=tg}, \ref{eq:tg=normal},
\ref{eq:normal=ambient} and Lemma \ref{l:diffvsconn},
\begin{equation}\label{eq:psi_*}
(\Psi_{\mathcal{C}}\circ\alpha)_*(v)\simeq \tnabla_v \alpha+v,
\end{equation}
where $\tnabla$ denotes the Levi-Civita connection on $T^*\mathcal{C}$.
\subsection{Third case: CS/AC Lagrangian submanifolds in $\C^m$}\label{ss:accslagdefs}
Let $\iota:L\rightarrow\C^m$ be an AC Lagrangian submanifold with rate
$\boldsymbol{\lambda}$, centers $p_i$ and ends $S_i$. Using the notation of Section
\ref{ss:conelagdefs},
the map $\Phi_{\mathcal{C}_i}+p_i:T^*\mathcal{C}_i\rightarrow\C^m$ identifies $\iota(S_i)\subset\C^m$ with the graph $\Gamma(\alpha_i)$ of some closed 1-form $\alpha_i$. This construction also determines a distinguished coordinate
system $\phi_i$ by imposing the relation
\begin{equation*}
\phi_i:\mathcal{C}_i\rightarrow S_i, \ \ \iota\circ\phi_i=\Phi_{\mathcal{C}_i}\circ\alpha_i.
\end{equation*}
Letting $(d\phi_i)^*:T^*S_i\rightarrow T^*\mathcal{C}_i$
denote the corresponding identification of cotangent bundles, we obtain an
identification of the zero section $\mathcal{C}_i$ with the zero section $S_i$.
We can use the symplectomorphism $\tau_{\alpha_i}$ defined in Equation
\ref{eq:translation} to ``bridge the gap'' between these identifications,
obtaining a symplectomorphism
\begin{equation}
\Phi_{S_i}:\mathcal{U}_i\subset T^*S_i\rightarrow \C^m, \ \
\Phi_{S_i}:=\Phi_{\mathcal{C}_i}\circ\tau_{\alpha_i}\circ (d\phi_i)^*+p_i
\end{equation}
which restricts to the identity on $S_i$. These maps provide a Lagrangian
neighbourhood for each end of $L$. Using the same methods as in the proof of Theorem
\ref{th:nbd_weinstein} one can interpolate between these maps. The final result
is a symplectomorphism
\begin{equation}\label{eq:aclagmap}
\Phi_L:\mathcal{U}\subset T^*L\rightarrow \C^m
\end{equation}
which restricts to the identity along $L$. This allows us to parametrize AC
deformations of $L$ with rate $\boldsymbol{\lambda}$ in terms of closed 1-forms
in the space $C^\infty_{\boldsymbol{\lambda}-1}(\mathcal{U})$.
More generally, given a CS or CS/AC Lagrangian submanifold $L$ in $\C^m$, the same
ideas define a symplectomorphism $\Phi_{L}$ as in Equation \ref{eq:aclagmap}. The same is true for a CS
submanifold in $M$: this time it is necessary to insert appropriate compositions
by $\Upsilon_i$. We refer to Joyce \cite{joyce:I} for additional details
concerning constructions of this type.
Coupling these results with Decompositions \ref{decomp:closedforms_growth},
\ref{decomp:closedforms_decay} and \ref{decomp:closedforms_accs} now gives a
good idea of the local structure of the corresponding moduli spaces of
Lagrangian deformations, defined as in Sections \ref{ss:cptlagdefs} and
\ref{ss:conelagdefs}.
\subsection{Lagrangian deformations with moving
singularities}\label{ss:movingsings}
In Section \ref{ss:accslagdefs} the given Lagrangian submanifold $L$ is deformed
keeping the singular points fixed in the ambient manifold $\C^m$ or $M$. It is
also natural to want to deform $L$ allowing the singular points to move within
the ambient space. Analogously, one might want to allow the corresponding
Lagrangian cones $\cone_i$ to rotate in $\C^m$. The correct set-up for doing
this when $\iota:L\rightarrow M$ is a CS Lagrangian submanifold with singularities
$\{x_1,\dots,x_s\}$ and identifications $\upsilon_i$ is as follows. The ideas are based on \cite{joyce:II}
Section 5.1. Define
\begin{equation}
P:=\{(p,\upsilon):p\in M,\ \upsilon:\C^m\rightarrow T_pM \mbox{ such that
}\upsilon^*\omega=\tilde{\omega}, \upsilon^*J=\tilde{J}\}.
\end{equation}
$P$ is a $\unitary m$-principal fibre bundle over $M$ with the action
$$\unitary m\times P\rightarrow P,\ \ M\cdot (p,\upsilon):=(p,\upsilon\circ
M^{-1}).$$
As such, $P$ is a smooth manifold of dimension $m^2+2m$.
Our aim is to use one copy of $P$ to parametrize the location of each singular
point $p_i=\iota(x_i)\in M$ and the direction of the corresponding cone $\cone_i\subset
\C^m$: the group action will allow the cone to rotate leaving the singular point
fixed. As we are interested only in small deformations of $L$ we can restrict
our attention to a small open neighbourhood of the pair $(p_i,\upsilon_i)\in P$. In general the
$\cone_i$ will have some symmetry group $G_i\subset \unitary m$, \textit{i.e.}
the action of this $G_i$ will leave the cone fixed. To ensure that we have no
redundant parameters we must therefore further restrict our attention to a
\textit{slice} of our open neighbourhood, \textit{i.e.} a smooth submanifold
transverse to the orbits of $G_i$. We denote this slice $\E_i$: it is a subset
of $P$ containing $(p_i,\upsilon_i)$ and of dimension $m^2+2m-\mbox{dim}(G_i)$.
We then set $\E:=\E_1\times\dots\times\E_s$. The point
$e:=(p_1,\upsilon_1),\dots,(p_s,\upsilon_s))\in \E$ will denote the initial
data as in Definition \ref{def:generalcslagsub}.
We now want to extend the datum of $(L,\iota)$ to a family of Lagrangian submanifolds
$(L,\iota_{\tilde{e}})$ parametrized by
$\tilde{e}=((\tilde{p}_1,\tilde{\upsilon}_1),\dots,(\tilde{p}_s,\tilde{\upsilon}_s))\in
\E$ (making $\E$ smaller if necessary). Each $(L,\iota_{\tilde{e}})$ should satisfy $\iota_{\tilde{e}}(p_i)=\tilde{p_i}$ and admit identifications $\tilde{\upsilon}_i$ and cones $\mathcal{C}_i$
as in Definition \ref{def:generalcslagsub}. We further require that $\iota_e=\iota$ globally and that $\iota_{\tilde{e}}=\iota$ outside a neighbourhood of the singularities.
The construction of such a family is actually straight-forward: using the maps
$\Upsilon_i$, it reduces to a choice of an appropriate family of
compactly-supported symplectomorphisms of $\C^m$.
It is now possible to choose an open neighbourhood $\mathcal{U}\subset T^*L$ and
embeddings $\Phi_{L}^{\tilde{e}}:\mathcal{U}\rightarrow M$ which, away from the
singularities, coincide with the embedding $\Phi_{L}$ introduced in Section
\ref{ss:accslagdefs}. The final result is that, after such a choice, the
\textit{moduli space of CS Lagrangian deformations of $L$ with rate
$\boldsymbol{\mu}$ and moving singularities} can be parametrized in terms of
pairs $(\tilde{e}, \alpha)$ where $\tilde{e}\in\E$ and $\alpha$ is a closed 1-form
on $L$ belonging to the space $C^\infty_{\boldsymbol{\mu}-1}(\mathcal{U})$.
Analogous results hold of course for CS and CS/AC submanifolds in $\C^m$. In
this case it is sufficient to set $P:=\{(p,\upsilon)\}$, with $p\in\C^m$ and
$\upsilon\in \unitary m$.
\section{Special Lagrangian conifolds}\label{s:slgeometry}
\begin{definition}\label{def:cy} A \textit{Calabi-Yau} (CY) manifold is the data
of a K\"ahler manifold ($M^{2m}$,$g$,$J$,$\omega$) and a non-zero ($m,0$)-form
$\Omega$ satisfying $\nabla\Omega\equiv 0$ and normalized by the condition
$\omega^m/m!=(-1)^{m(m-1)/2}(i/2)^m\Omega\wedge\bar{\Omega}$.
In particular $\Omega$ is holomorphic and the holonomy of $(M,g)$ is contained
in $\sunitary m$. We will refer to $\Omega$ as the \textit{holomorphic volume
form} on $M$.
\end{definition}
\begin{example}\label{e:C^m}
The simplest example of a CY manifold is $\C^m$ with its standard structures
$\tilde{g}$, $\tilde{J}$, $\tilde{\omega}$ and
$\tilde{\Omega}:=dz^1\wedge\dots\wedge dz^m$.
\end{example}
\begin{definition}\label{def:sl}
Let $M^{2m}$ be a CY manifold and $L^m\rightarrow M$ be an immersed or embedded
Lagrangian submanifold. We can restrict $\Omega$ to $L$, obtaining a
non-vanishing complex-valued $m$-form $\Omega_{|L}$ on $L$. We say that $L$ is
\textit{special Lagrangian} (SL) iff this form is real, \textit{i.e.}
$\Imag\,\Omega_{|L}\equiv 0$. In this case $\Real\,\Omega_{|L}$ defines a volume
form on $L$, thus a natural orientation.
\end{definition}
Lagrangian submanifolds (especially the immersed ones) tend to be very ``soft"
objects: for example, Section \ref{s:lagdefs} shows that they have
infinite-dimensional moduli spaces. They also easily allow for cutting, pasting
and desingularization procedures. The ``special" condition rigidifies them
considerably: the corresponding deformation, gluing and desingularization
processes require much ``harder" techniques. Cf. \textit{e.g.}
\cite{haskinskapouleas}, \cite{joyce:III}, \cite{joyce:IV}, \cite{pacini:slgluing} for recent gluing
results and \cite{haskinspacini} for local desingularization issues.
\begin{definition} \label{def:accssl}
We can define AC, CS and CS/AC special Lagrangian submanifolds in $\C^m$ exactly
as in Definitions \ref{def:aclagsub}, \ref{def:cslagsub} and
\ref{def:accslagsub}, simply adding the requirement that the submanifolds be
special Lagrangian. In particular this implies that the cones $\mathcal{C}_i$
are SL in $\C^m$. Following Definition \ref{def:generalcslagsub} we can also
define CS special Lagrangian submanifolds in a general CY manifold $M$: in this
case it is necessary to also add the requirement that
$\upsilon_i^*\Omega=\tilde{\Omega}$.
We use the generic term \textit{special Lagrangian conifold} to refer to any of
the above.
\end{definition}
\begin{remark} \label{rem:muregularity}
It follows from Joyce \cite{joyce:I} Theorem 5.5 that if $L$ is a CS or CS/AC SL
submanifold with respect to some rate $\boldsymbol{\mu}=2+\epsilon$ with
$\epsilon$ in a certain range $(0,\epsilon_0)$ then it is also CS or CS/AC with
respect to any other rate of the form $\boldsymbol{\mu}'=2+\epsilon'$ with
$\epsilon'\in (0,\epsilon_0)$. The precise value of $\epsilon_0$ is determined
by certain \textit{exceptional weights} for the cones $\mathcal{C}_i$,
introduced in Section \ref{s:reviewlaplace}. We refer to \cite{joyce:I} for
details.
\end{remark}
\begin{example} \label{e:slcone}
Let $\mathcal{C}$ be a Lagrangian cone in $\C^m$ with smooth link
$\Sigma^{m-1}$. It can be shown that $\mathcal{C}$ is SL (with respect to some
holomorphic volume form $e^{i\theta}\tilde{\Omega}$) iff $\Sigma$ is minimal in
$\Sph^{2m-1}$ with respect to the natural metric on the sphere. Then
$\mathcal{C}$ is a CS/AC SL in $\C^m$. Cf. \textit{e.g.} \cite{harveylawson},
\cite{haskins}, \cite{haskinskapouleas}, \cite{haskinspacini}, \cite{joyce:symmetries} for examples.
We refer to Joyce \cite{joyce:V} Section 6.4 for examples of AC SLs in $\C^m$
with various rates.
\end{example}
\section{Setting up the SL deformation problem}\label{s:setup}
If $\iota:L\rightarrow M$ is a SL conifold we can specialize the framework of Section
\ref{s:lagdefs} to study the SL deformations of $L$. Notice that the SL
condition is again invariant under reparametrizations. Thus, if $L$ is smooth
and compact, the \textit{moduli space} $\mathcal{M}_L$ \textit{of SL
deformations of $L$} can be defined as the connected component containing $L$ of
the subset of SL submanifolds in $\mbox{Lag}(L,M)/\mbox{Diff}(L)$. As seen in
Sections \ref{ss:conelagdefs} and \ref{ss:accslagdefs}, if $L$ is an AC, CS or
CS/AC Lagrangian submanifold with specific rates of growth/decay on the ends, we can obtain
moduli spaces of Lagrangian or SL deformations of $L$ with those same rates by
simply restricting our attention to closed 1-forms on $L$ which satisfy
corresponding growth/decay conditions.
Our ultimate goal is to prove that moduli spaces of SL conifolds often admit a
natural smooth structure with respect to which they are finite-dimensional
manifolds. Failing this, we want to identify the obstructions which prevent this
from happening. Generally speaking, the strategy for proving these results will
be to view $\mathcal{M}_L$ locally as the zero set of some smooth map $F$
defined on the space of closed forms in $C^\infty(\mathcal{U})$ (when $L$ is
smooth and compact) or in
$C^\infty_{(\boldsymbol{\mu}-1,\boldsymbol{\lambda}-1)}(\mathcal{U})$ (when $L$
is CS/AC with rate $(\boldsymbol{\mu},\boldsymbol{\lambda})$): we can then
attempt to use the Implicit Function Theorem to prove that this zero set
is smooth.
The choice of $F$ is dictated by Definition \ref{def:sl}. Let $\Omega$ denote
the given holomorphic volume form on $M$. Then $F$ must compute the values
of $\Imag\,\Omega$ on each Lagrangian deformation of $L$. In the following
sections we present the precise construction of $F$ and study its properties,
for each case of interest.
\textit{Note: }To simplify the notation, from now on we will drop the immersion $\iota:L\rightarrow M$ and simply identify $L$ with its image. In particular we will identify the singularities $x_i$ with their images $\iota(x_i)$.
\subsection{First case: smooth compact special Lagrangians}\label{ss:cptslsetup}
Let $L\subset M$ be a smooth compact SL submanifold, endowed with the induced
metric $g$ and orientation. Define $\Phi_L:\mathcal{U}\rightarrow M$ as in
Section \ref{ss:cptlagdefs}. Consider the pull-back real $m$-form
$\Phi_L^*(\Imag\,\Omega)$ defined on $\mathcal{U}$. Given any closed $\alpha\in
C^\infty(\mathcal{U})$, let $\Gamma(\alpha)$ denote the submanifold in
$\mathcal{U}$ defined by its graph. It is diffeomorphic to $L$ via the
projection $\pi:T^*L\rightarrow L$. The pull-back form restricts to an $m$-form
$\Phi_L^*(\Imag\,\Omega)_{|\Gamma(\alpha)}$ on $\Gamma(\alpha)$. It is clear
from Definition \ref{def:sl} that $\Gamma(\alpha)$ is SL iff this form vanishes.
We can now pull this form back to $L$ via $\alpha$ (equivalently, push it down to $L$ via $\pi_*$), obtaining a real $m$-form on
$L$: then $\Gamma(\alpha)$ is SL iff this
form vanishes on $L$. Finally, let $\star$ denote the \textit{Hodge
star operator} defined on $L$ by $g$ and the orientation. Using this operator we
can reduce any $m$-form on $L$ to a function.
Summarizing, let $\mathcal{D}_L$ denote the space of closed 1-forms on $L$ whose graph lies
in $\mathcal{U}$. We then define the map $F$ as follows.
\begin{equation}\label{eq:defF}
F:\mathcal{D}_L\rightarrow C^\infty (L),\ \ \alpha\mapsto
\star(\alpha^*(\Phi_L^*\Imag\,\Omega))=\star((\Phi_L\circ\alpha)^*\Imag\,\Omega).
\end{equation}
\begin{prop} \label{prop:cptnonlinear}
The non-linear map $F$ has the following properties:
\begin{enumerate}
\item The set $F^{-1}(0)$ parametrizes the space of all SL deformations of $L$
which are $C^1$-close to $L$.
\item $F$ is a smooth map between Fr\'echet spaces. Furthermore, for each
$\alpha\in\mathcal{D}_L$, $\int_LF(\alpha)\,vol_g=0$.
\item The linearization $dF[0]$ of $F$ at $0$ coincides with the operator $d^*$,
\textit{i.e.}
\begin{equation}\label{eq:linearization}
dF[0](\alpha)=d^*\alpha.
\end{equation}
\end{enumerate}
\end{prop}
\begin{proof} These results are standard, cf. \cite{mclean} or \cite{joyce:I}
Prop. 2.10. However for the reader's convenience we give a sketch of the
argument with respect to our own set of conventions. To simplify the notation we
identify $\mathcal{U}$ with its image in $M$ via $\Phi_L$. This allows us to
write
\begin{equation}
F(\alpha)=\star(\pi_*(\Imag\,\Omega_{|\Gamma(\alpha)})).
\end{equation}
We also identify $L$ with the zero section in $T^*L$.
The first statement follows directly from the definition of $F$ and the results
of Section \ref{ss:cptlagdefs}. More precisely the statement is that, up to
composition with $\Phi_L$, $F^{-1}(0)$ coincides with the set of SL submanifolds
which admit a parametrization which is $C^1$-close to some parametrization of
$L$.
To prove the second statement, notice that
$\int_LF(\alpha)\,vol_g=\int_{\Gamma(\alpha)}\Imag\,\Omega$. The fact that
$\Omega$ is closed implies that $\Imag\,\Omega$ is closed. Furthermore the
submanifold $\Gamma(\alpha)$ is homotopic, thus homologous, to the zero section
$L$. Thus $\int_{\Gamma(\alpha)}\Imag\,\Omega=\int_L\Imag\,\Omega=0$ because $L$
is SL. The smoothness of $F$ is clear from its definition.
To prove Equation \ref{eq:linearization}, fix any $\alpha\in \Lambda^1(L)$ and
let $v$ denote the normal vector field along $L$ determined by imposing
$\alpha(\cdot)\equiv\omega(v,\cdot)$. We can extend $v$ to a global vector field
$v$ on $M$. Let $\phi_s$ denote any 1-parameter family of diffeomorphisms of $M$
such that $d/ds(\phi_s(x))_{|s=0}=v(x)$. Then the two 1-parameter families of
$m$-forms on $L$, $(s\alpha)^*(\Imag\,\Omega)=\pi_*(\Imag \,\Omega_{|\Gamma(s\alpha)})$ and
$(\phi_s^*\Imag\,\Omega)_{|L}$, coincide up to first order so that standard
calculus of Lie derivatives shows that
\begin{eqnarray*}
dF[0](\alpha)\,vol_g &=& d/ds
(F(s\alpha)\,vol_g)_{|s=0}\\
&=& d/ds(\phi_s^*\Imag\,\Omega)_{|L;\,s=0}\\
&=& (\mathcal{L}_v\Imag\,\Omega)_{|L}= (di_v\Imag\,\Omega)_{|L},
\end{eqnarray*}
where in the last equality we use \textit{Cartan's formula}
$\mathcal{L}_v=di_v+i_vd$ and the fact that $\Imag\,\Omega$ is closed.
We now claim that $(i_v\Imag\,\Omega)_{|L}\equiv-\star\alpha$ on $L$. This is a
linear algebra statement so we can check it point by point. We can also assume
that $v$ is a unit vector at that point. Fix a point $x\in L$ and an isomorphism
$T_xM\simeq \C^m$ identifying the CY structures on $T_xM$ with the standard
structures on $\C^m$. This map will identify $T_xL$ with a SL $m$-plane $\Pi$ in
$\C^m$. Consider the action of $\sunitary m$ on the Grassmannian of $m$-planes
in $\C^m$. In \cite{harveylawson} page 89 it is shown that $\sunitary m$ acts
transitively on the subset of SL $m$-planes and that the isotropy subgroup
corresponding to the distinguished SL plane $\R^m:=\mbox{span}\{\partial
x^1,\dots,\partial x^m\}$ is $\sorth m\subset \sunitary m$; in other words, the
set of SL $m$-planes in $\C^m$ can be identified with the homogeneous space
$\sunitary m/\sorth m$. Up to a rotation in $\sunitary m$ we can assume that
$\Pi=\R^m$. Up to a rotation in $\sorth m$ we can further assume that
$v(x)=\partial y^1$. It is thus sufficient to check our claim in this case only.
We can write $\Imag\,\Omega=dy^1\wedge dx^2\wedge\dots\wedge dx^m+(\dots)$. It
follows that $(i_v\Imag\,\Omega)_{|\R^m}=dx^2\wedge\dots\wedge dx^m$. On the
other hand $\alpha=-dx^1$, proving the claim, thus Equation
\ref{eq:linearization}.
\end{proof}
\begin{remark} \label{rem:cptQ}
Notice that $\star$ depends on $x\in L$, $\Gamma(\alpha)$ depends on $\alpha$
and $\Phi_L^*\Imag\,\Omega_{|\Gamma(\alpha)}$ depends on $\alpha$ and
$\nabla\alpha$. We can thus think of $F$ as being obtained from an underlying
smooth function
\begin{equation}
F'=F'(x,y,z):\mathcal{U}\oplus(T^*L\otimes T^*L)\rightarrow \R
\end{equation}
via the following relationship:
\begin{equation}\label{eq:F_unwrapped}
F(\alpha)=F'(x,\alpha(x),\nabla\alpha(x)).
\end{equation}
More specifically, $F'$ can be defined as follows. Choose a point $(x,y)\in
\mathcal{U}$. Let $e_1,\dots,e_m$ be an orthonormal positive basis of $T_xL$.
Now choose any $z\in T_x^*L\otimes T_x^*L$. Recall from Lemma \ref{l:diffvsconn}
that, using the Levi-Civita connection, $T_{(x,y)}\mathcal{U}\simeq T_x^*L\oplus
T_xL$. Thus the vectors $(i_{e_i}z,e_i)$ span an $m$-plane in
$T_{(x,y)}\mathcal{U}$; when $y=\alpha$ and $z=\nabla\alpha$, this $m$-plane
coincides with $T_{(x,\alpha)}\Gamma(\alpha)$. We can now define
\begin{equation}\label{eq:F'}
F'(x,y,z):=\Phi_L^*\Imag\,\Omega_{|(x,y)}((i_{e_1}z,e_1),\dots,(i_{e_m}z,
e_m)).
\end{equation}
For any fixed $x\in L$, $y$ and $z$ vary in the linear space
$T^*_xL\oplus(T^*_xL\otimes T^*_xL)$ so Taylor's theorem shows
\begin{equation}
F'(x,y,z)=F'(x,0,0)+\frac{\partial F'}{\partial
y}(x,0,0)\,y+\frac{\partial F'}{\partial z}(x,0,0)\,z+Q'(x,y,z)
\end{equation}
for some smooth $Q'=Q'(x,y,z)$ satisfying
$Q'(x,y,z)=O(|y|^2+|z|^2)$ for each $x$, as $|y|\rightarrow 0$ and
$|z|\rightarrow 0$. By substitution we find
\begin{eqnarray*}
F(\alpha) &=& F'(x,\alpha(x),\nabla\alpha(x))\\
&=&F'(x,0,0)+\frac{\partial F'}{\partial
y}(x,0,0)\,\alpha(x)+\frac{\partial F'}{\partial
z}(x,0,0)\,\nabla\alpha(x)+Q'(x,\alpha(x),\nabla\alpha(x)).
\end{eqnarray*}
The fact that $L$ is SL implies that $F'(x,0,0)\equiv 0$. Notice also that
by the chain rule
\begin{equation*}
d/ds(F(s\alpha))_{|s=0}=d/ds(F'(x,s\alpha(x),s\nabla\alpha(x))_{|s=0}=\frac
{\partial F'}{\partial y}(x,0,0)\,\alpha(x)+\frac{\partial
F'}{\partial z}(x,0,0)\,\nabla\alpha(x).
\end{equation*}
On the other hand, $d/ds(F(s\alpha))_{|s=0}=dF[0](\alpha)=d^*\alpha$.
Combining these equations leads to
\begin{equation}\label{eq:cptQ}
F(\alpha)=d^*\alpha+Q'(x,\alpha(x),\nabla\alpha(x)).
\end{equation}
\end{remark}
\subsection{Second case: special Lagrangian cones in
$\C^m$}\label{ss:coneslsetup}
Let $\mathcal{C}$ be a SL cone in $\C^m$, endowed with the induced metric
$\tilde{g}$ and orientation. Define $\Phi_{\mathcal{C}}:\mathcal{U}\rightarrow
\C^m$ as in Section \ref{ss:conelagdefs}. Fix any $\mu>2$, $\lambda<2$. Let
$\mathcal{D}_{\mathcal{C}}$ denote the space of closed 1-forms in
$C^\infty_{(\mu-1,\lambda-1)}(\Lambda^1)$ whose graph lies in $\mathcal{U}$.
Given $\alpha\in \mathcal{D}_{\mathcal{C}}$, define $F(\alpha)$ as in Equation
\ref{eq:defF}.
\begin{prop} \label{prop:conenonlinear}
The non-linear map $F$ has the following properties:
\begin{enumerate}
\item The set $F^{-1}(0)$ parametrizes the space of all SL deformations of
$\mathcal{C}$ which are $C^1$-close to $L$ and are asymptotic to $\mathcal{C}$
with rate $(\mu,\lambda)$.
\item $F$ is a well-defined smooth map
\begin{equation*}
F:\mathcal{D}_{\mathcal{C}}\rightarrow
C^\infty_{(\mu-2,\lambda-2)}(\mathcal{C}).
\end{equation*}
In particular, for each $\alpha\in\mathcal{D}_{\mathcal{C}}$, $F(\alpha)\in
C^\infty_{(\mu-2,\lambda-2)}(\mathcal{C})$.
\item The linearization $dF[0]$ of $F$ at $0$ coincides with the operator $d^*$,
\textit{i.e.}
\begin{equation}
dF[0](\alpha)=d^*\alpha.
\end{equation}
\end{enumerate}
\end{prop}
\begin{proof}
The first statement follows from the definition of $F$ and the results of
Section \ref{ss:conelagdefs}. Concerning the second statement, we may write
\begin{eqnarray*}
F(\alpha) &=&
\star(\alpha^*(\Phi_{\mathcal{C}}^*\Imag\,\tilde{\Omega}))=\Imag\,\tilde{\Omega}((\Phi_{\mathcal{C}}\circ\alpha)_*(e_1),\dots,(\Phi_{
\mathcal{C}}\circ\alpha)_*(e_m))\\
&=&
\Imag\,\tilde{\Omega}((\Psi_{\mathcal{C}}
\circ\alpha)_*(e_1)+(R\circ\alpha)_*(e_1),\dots,(\Psi_{\mathcal{C}}
\circ\alpha)_*(e_m)+(R\circ\alpha)_*(e_m))\\
&=&
\Imag\,\tilde{\Omega}((\Psi_{\mathcal{C}}\circ\alpha)_*(e_1),\dots,(\Psi_{
\mathcal{C}}\circ\alpha)_*(e_m))+\dots,\\
\end{eqnarray*}
where $e_i$ is a local $\tg$-orthornomal basis of $T\mathcal{C}$.
Consider this last equation as $r\rightarrow\infty$. Equation \ref{eq:psi_*}
shows that its first term is of the form
$\Imag\,\tilde{\Omega}(e_1,\dots,e_m)+O(r^{\lambda-2})$. The first term here
vanishes because $\mathcal{C}$ is SL, leaving the term $O(r^{\lambda-2})$.
Equation \ref{eq:Rscaling} shows that the remaining terms in $F(\alpha)$ are of
the form $O(r^{2\lambda-4})$. Analogous methods apply for $r\rightarrow 0$,
showing that $F(\alpha)\in C^0_{(\mu-2,\lambda-2)}(\mathcal{C})$.
To study the derivatives of $F(\alpha)$ we endow $\mathcal{U}$ with the metric and Levi-Civita connection $\nabla$ pulled back from $\C^m$ via $\Phi_{\mathcal{C}}$, so that $\nabla(\Phi_{\mathcal{C}}^*\Imag\,\tilde{\Omega})=\Phi_{\mathcal{C}}^*(\tnabla\Imag\,\tilde{\Omega})=0$. Let $g$ denote the induced metric on $\Gamma(\alpha)$. Then $\mathcal{C}$ can be endowed with either the metric $\tg$ and induced connection $\tnabla$ or with the metric $\alpha^*g$ and induced connection $\nabla$. One can check, or cf. \cite{pacini:weighted}, that the fact that $\alpha^*g$ is asymptotic to $\tg$ implies that the difference tensor $A:=\nabla-\tnabla$ satisfies $|A|=O(r^{\lambda-3})$, as $r\rightarrow \infty$. Notice that
\begin{equation*}
F(\alpha)\,\mbox{vol}_{\tilde{g}}=(\Phi_{\mathcal{C}}\circ\alpha)^*\Imag\,\tilde{\Omega}
\end{equation*}
so, taking derivatives,
\begin{equation*}
\nabla(F(\alpha)\,\mbox{vol}_{\tilde{g}})=\nabla((\Phi_{\mathcal{C}}\circ\alpha)^*\Imag\,\tilde{\Omega})=(\Phi_{\mathcal{C}}\circ\alpha)^*(\tnabla\Imag\,\tilde{\Omega})=0.
\end{equation*}
This implies
\begin{equation*}
|(\nabla F(\alpha))\otimes\mbox{vol}_{\tg}|=|F(\alpha)\cdot\nabla(\mbox{vol}_{\tg})|=O(r^{\lambda-2})|\nabla(\mbox{vol}_{\tg})|.
\end{equation*}
Write $\mbox{vol}_{\tg}=e_1^*\otimes\dots\otimes e_m^*$ so that $\nabla(\mbox{vol}_{\tg})=\nabla e_1^*\otimes\dots\otimes e_m^*+\dots+e_1^*\otimes\dots\otimes\nabla e_m^*$. We may assume that $\tnabla e_i^*=0$. Then $\nabla e_i^*=(\nabla-\tnabla) e_i^*=A e_i^*$, leading to $|\nabla(\mbox{vol}_{\tg})|=O(r^{\lambda-3})$.
This shows that $F(\alpha)\in C^1_{(\mu-2,\lambda-2)}(\mathcal{C})$.
Further
calculations of the same type apply to the higher derivatives, showing that
$F(\alpha)\in C^\infty_{(\mu-2,\lambda-2)}(\mathcal{C})$. It is clear that $F$
is smooth.
The third statement can be proved as in Proposition \ref{prop:cptnonlinear}.
\end{proof}
\subsection{Third case: CS/AC special Lagrangians in $\C^m$}\label{ss:accsslsetup}
Let $L$ be a AC, CS or CS/AC SL in $\C^m$ or a CS SL in $M$. The moduli space of
SL deformations of $L$ with fixed singularities coincides locally with the zero
set of a map $F$ defined as in Equation \ref{eq:defF}. The methods and results
of Sections \ref{ss:accslagdefs} and \ref{ss:coneslsetup} then lead to a good
understanding of the properties of $F$, analogous to those described in
Propositions \ref{prop:cptnonlinear} and \ref{prop:conenonlinear}. For example,
assume $L$ is a CS/AC SL in $\C^m$ with rate
$(\boldsymbol{\mu},\boldsymbol{\lambda})$. Let $\mathcal{D}_{L}$ denote the
space of closed 1-forms in
$C^\infty_{(\boldsymbol{\mu}-1,\boldsymbol{\lambda}-1)}(\Lambda^1)$ whose graph
lies in $\mathcal{U}$. Choose $\alpha\in \mathcal{D}_{L}$. Then it is simple to
check that $F(\alpha)\in
C^\infty_{(\boldsymbol{\mu}-2,\boldsymbol{\lambda}-2)}(L)$, \textit{etc}.
We now want to understand how to parametrize the SL deformations of $L$ whose
singularities are allowed to move in the ambient space as in Section
\ref{ss:movingsings}. For example, assume $L$ is a CS SL submanifold in $M$. The
constructions of Section \ref{ss:movingsings} must then be modified as follows.
This time we set
\begin{equation}
\tilde{P}:=\{(p,\upsilon):p\in M,\ \upsilon:\C^m\rightarrow T_pM \mbox{ such
that }\upsilon^*\omega=\tilde{\omega},\ \upsilon^*\Omega=\tilde{\Omega}\},
\end{equation}
so that $\tilde{P}$ is a $\sunitary m$-principal fibre bundle over $M$ of
dimension $m^2+2m-1$. For each end, the cone $\cone_i$ will now have symmetry
group $G_i\subset \sunitary m$. As in Section \ref{ss:movingsings}, let
$\tilde{\E}_i$ denote a smooth submanifold of $\tilde{P}$ transverse to the
orbits of $G_i$. It has dimension $m^2+2m-1-\mbox{dim}(G_i)$. Set
$\tilde{\E}:=\tilde{\E}_1\times\dots\times\tilde{\E}_s$. We then define CS
Lagrangian submanifolds $L_{\tilde{e}}$ and embeddings $\Phi_{L}^{\tilde{e}}$ with
the same properties as before.
Now let $\mathcal{D}_{L}$ denote the space of closed 1-forms in
$C^\infty_{\boldsymbol{\mu}-1}(\Lambda^1)$ whose graph lies in $\mathcal{U}$. We
define a map
\begin{equation}\label{eq:csdefF}
F:\tilde{\E}\times\mathcal{D}_{L}\rightarrow C^\infty_{\boldsymbol{\mu}-2} (L),\
\ (\tilde{e},\alpha)\mapsto
\star(\alpha^*(\Phi_{L}^{\tilde{e}*}\Imag\,\Omega)).
\end{equation}
\begin{prop} \label{prop:csnonlinear}
Let $L$ be a CS SL in $M$. Then the map $F$ has the following properties:
\begin{enumerate}
\item The set $F^{-1}(0)$ parametrizes the space of all SL deformations of $L$
which are $C^1$-close to $L$ away from the singularities and are asymptotic to
$\mathcal{C}_i$ with rate $\mu_i$ for some choice of
$(\tilde{p}_i,\tilde{\upsilon}_i)$ near $(p_i,\upsilon_i)$.
\item $F$ is a (locally) well-defined smooth map between Fr\'echet spaces. In
particular, for each $\alpha\in\mathcal{D}_{L}$, $F(\alpha)\in
C^\infty_{\boldsymbol{\mu}-2}(L)$. Furthermore, $\int_L F(\alpha)\,vol_g=0$.
\item There exists an injective linear map $\chi:T_e\tilde{\E}\rightarrow
C^\infty_{\boldsymbol{0}}(L)$ such that (i) $\chi(y)\equiv 0$ away from the
singularities and (ii) the linearized map $dF[0]:T_e\tilde{\E}\oplus
C^\infty_{\boldsymbol{\mu}-1}(\Lambda^1)\rightarrow
C^\infty_{\boldsymbol{\mu}-2}(L)$ satisfies
\begin{equation}\label{eq:cslinearization}
dF[0](y,\alpha)=\Delta_g\,\chi(y)+d^*\alpha.
\end{equation}
\end{enumerate}
\end{prop}
\begin{proof}
The first statement should be interpreted as explained in the proof of
Proposition \ref{prop:cptnonlinear}. The proof follows from the definitions of
$\tilde{\E}$ and $F$ and from the results of Section \ref{ss:movingsings}. The
second statement can be proved as in Propositions \ref{prop:cptnonlinear} and
\ref{prop:conenonlinear}.
Regarding the third statement, the linearization of $F$ with respect to
directions in $C^\infty_{\boldsymbol{\mu}-1}(\Lambda^1)$ can be computed as in
Proposition \ref{prop:cptnonlinear}. Now choose $y\in T_e\tilde{\E}$
corresponding to a curve $\tilde{e}_s\in\tilde{\E}$ such that $\tilde{e}_0=e$. Up to
identifying $\mathcal{U}$ with $M$ via $\Phi_{L}^e$, $\Phi_{L}^{\tilde{e}_s}$
defines a 1-parameter curve of symplectomorphisms $\phi_s$ of $M$ such that
$d/ds(\phi_s)_{|s=0}=v$, for some vector field $v$ on $M$. Thus, as in
Proposition \ref{prop:cptnonlinear},
\begin{eqnarray*}
dF[0](y)\,vol_g &=&
d/ds(F(\tilde{e}_s,0)\,vol_g)_{|s=0}=d/ds((\phi_s)^*\Imag\,\Omega)_{|L;s=0}\\
&=& (\mathcal{L}_v\Imag\,\Omega)_{L}=(di_v\Imag\Omega)_{|L}\\
&=&-d\star\alpha,
\end{eqnarray*}
where $\alpha:=\omega(v,\cdot)_{|L}$ is a closed 1-form on $L$. Notice that, by
definition, $\phi_s\equiv Id$ away from the singularities of $L$, so
$\alpha\equiv 0$ there. Thus, by the Poincar\'e Lemma (cf. \textit{e.g.} Lemma
\ref{lemma:formclosed}), $\alpha$ must be exact on $L$, \textit{i.e.}
$\alpha=d\chi$ for some function $\chi:L\rightarrow \R$. We can define $\chi$
uniquely by imposing that $\chi\equiv 0$ away from the singularities of $L$. The
function $\chi$ depends linearly on $y$, and we can write
$dF[0](y,0)=\Delta_g\,\chi(y)$, as claimed. Furthermore, if $\chi(y)=0$ then
$\alpha=0$ and $v=0$. Since $\tilde{\E}$ is defined so as to parametrize
geometrically distinct immersions, this implies $y=0$.
Roughly speaking, near each singularity and up to the appropriate
identifications, $\tilde{e}_s$ should be thought of as a 1-parameter curve in the
group $\sunitary m\ltimes\C^m$ acting on $\C^m$. This action admits a
\textit{moment map} $\mu:\C^m\rightarrow (Lie(\sunitary m\ltimes\C^m))^*$.
Recall that this means that $\mu$ is equivariant and that, for all $w\in
Lie(\sunitary m\ltimes\C^m)$, the corresponding function
$\mu_w:\C^m\rightarrow\R$ satisfies $d\mu_w=i_w\tilde{\omega}$, \textit{i.e.}
$w$ is a Hamiltonian vector field with Hamiltonian function $\mu_w$. The moment
map can be written explicitly, cf. \textit{e.g.} \cite{haskinspacini} Section
2.6, showing that each $\mu_w$ is at most a quadratic polynomial on $\C^m$.
Notice, for future reference, that for any SL cone $\mathcal{C}\subset\C^m$ the
calculations in the proof of Proposition \ref{prop:cptnonlinear} show that
\begin{equation*}
\Delta_g(\mu_{w|\mathcal{C}})=d^*(d\mu_{w|\mathcal{C}})=-\star
d\star(i_w\tilde{\omega}_{|\mathcal{C}})=\star (d
i_w\Imag\,\tilde{\Omega})_{|\mathcal{C}}=\star(\mathcal{L}_w\Imag\,\tilde{\Omega
})_{|\mathcal{C}}=0,
\end{equation*}
\textit{i.e.} each $\mu_w$ restricts to a harmonic function on each SL cone.
In this set-up our vector field $v$ is (locally) an element of $Lie(\sunitary
m\ltimes\C^m)$ and $\chi(y)=\mu_v$. Thus $\chi(y)$ is bounded as $r\rightarrow
0$. This implies that $\chi(y)\in C^0_{\boldsymbol{0}}(L)$. Further calculations
show that $\chi(y)\in C^\infty_{\boldsymbol{0}}(L)$, as claimed.
\end{proof}
\begin{remark}\label{rem:chidecay}
Recall from Section \ref{s:lagconifolds} that, on each conically singular end, the metric $g_i$ decays to the metric $\tilde{g}_i$ with rate
$\mu_i-2$. As shown in \cite{pacini:weighted}, this implies that
\begin{equation*}
|\Delta_{g_i}\,\chi(y)|=|\Delta_{\tg_i}\,\chi(y)|(1+O(r^{\mu_i-2}))+|d\,\chi(y)|_{g_i}O(r^{\mu_i-3}).
\end{equation*}
Using the fact that $\chi(y)$ is harmonic with respect to the cone metric, we see that $\Delta_g\,\chi(y)\in C^\infty_{\boldsymbol{\mu}-4}\subset C^\infty_{\boldsymbol{\mu}-2}$, as expected from Equation \ref{eq:csdefF}.
\end{remark}
Now let $L$ be a CS/AC SL in $\C^m$. Define $\tilde{P}$, $\tilde{\E}$,
\textit{etc.} analogously to the above (cf. Section \ref{ss:movingsings} for the
necessary modifications for the ambient space $\C^m$). Let $\mathcal{D}_{L}$
denote the space of closed 1-forms in
$C^\infty_{(\boldsymbol{\mu}-1,\boldsymbol{\lambda}-1)}(\Lambda^1)$ whose graph
lies in $\mathcal{U}$. Define $F$ as in Equation \ref{eq:csdefF}. The following
result can then be proved similarly to Proposition \ref{prop:csnonlinear}.
\begin{prop}\label{prop:accsnonlinear}
Let $L$ be a CS/AC SL in $\C^m$. Then the map $F$ has the following properties:
\begin{enumerate}
\item The set $F^{-1}(0)$ parametrizes the space of all SL deformations of $L$
which are $C^1$-close to $L$ away from the singularities and are asymptotic to
$\mathcal{C}_i$ with rate $(\boldsymbol{\mu},\boldsymbol{\lambda})$ for some
choice of $(\tilde{p}_i,\tilde{\upsilon}_i)$ near $(p_i,\upsilon_i)$.
\item $F$ is a (locally) well-defined smooth map between Fr\'echet spaces. In
particular, for each $\alpha\in\mathcal{D}_{L}$, $F(\alpha)\in
C^\infty_{(\boldsymbol{\mu}-2,\boldsymbol{\lambda}-2)}(L)$.
\item There exists an injective linear map $\chi:T_e\tilde{\E}\rightarrow
C^\infty_{\boldsymbol{0}}(L)$ such that (i) $\chi(y)\equiv 0$ away from the
singularities and (ii) the linearized map $dF[0]:T_e\tilde{\E}\oplus
C^\infty_{(\boldsymbol{\mu}-1,\boldsymbol{\lambda}-1)}(\Lambda^1)\rightarrow
C^\infty_{(\boldsymbol{\mu}-2,\boldsymbol{\lambda}-2)}(L)$ satisfies
\begin{equation}
dF[0](y,\alpha)=\Delta_g\,\chi(y)+d^*\alpha.
\end{equation}
\end{enumerate}
\end{prop}
If the spaces $C^\infty(L)$, $C^\infty_{\boldsymbol{\beta}}(L)$ were Banach
spaces and the relevant maps were Fredholm, we could now apply the Implicit
Function Theorem to conclude that the sets $F^{-1}(0)$, and thus
$\mathcal{M}_L$, are smooth. As however they are actually only Fr\'echet spaces,
it is instead necessary to first take the Sobolev space completions of these
spaces, then study the Fredholm properties of the linearized maps. We do this in
Section \ref{s:moduli}. This will require some results concerning the Laplace
operator on conifolds, summarized in Section \ref{s:reviewlaplace}.
\section{Review of the Laplace operator on conifolds}\label{s:reviewlaplace}
We summarize here some analytic results concerning the Laplace operator on
conifolds, referring to \cite{pacini:weighted} for further details and
references.
\begin{definition}\label{def:exceptionalweights}
Let $(\Sigma,g')$ be a compact Riemannian manifold. Consider the cone
$C:=\Sigma\times (0,\infty)$ endowed with the conical metric
$\tilde{g}:=dr^2+r^2g'$. Let $\Delta_{\tilde{g}}$ denote the corresponding
Laplace operator acting on functions.
For each component $(\Sigma_j,g_j')$ of $(\Sigma,g')$ and each $\gamma\in\R$,
consider the space of homogeneous harmonic functions
\begin{equation}\label{eq:ac_harmonicter}
V^j_{\gamma}:=\{r^\gamma\sigma(\theta):
\Delta_{\tilde{g}}(r^{\gamma}\sigma)=0\}.
\end{equation}
Set $m^j(\gamma):=\mbox{dim}(V^j_\gamma)$. One can show that $m^j_{\gamma}>0$
iff $\gamma$ satisfies the equation
\begin{equation}\label{eq:exceptionalforlaplacian}
\gamma=\frac{(2-m)\pm\sqrt{(2-m)^2+4e_n^j}}{2},
\end{equation}
for some eigenvalue $e_n^j$ of $\Delta_{g_j'}$ on $\Sigma_j$. Given any weight
$\boldsymbol{\gamma}\in \R^e$, we now set $m(\boldsymbol{\gamma}):=\sum_{j=1}^e
m^j(\gamma_j)$. Let $\mathcal{D}\subseteq\R^e$ denote the set of weights
$\boldsymbol{\gamma}$ for which $m(\boldsymbol{\gamma})>0$. We call these the
\textit{exceptional weights} of $\Delta_{\tg}$.
\end{definition}
Let $(L,g)$ be a conifold. Assume $(L,g)$ is asymptotic to a cone $(C,\tg)$ in
the sense of Definition \ref{def:conifold}. Roughly speaking, the fact that $g$
is asymptotic to $\tilde{g}$ in the sense of Definition \ref{def:metrics_ends}
implies that the Laplace operator $\Delta_g$ is ``asymptotic" to $\Delta_{\tg}$.
Applying Definition \ref{def:exceptionalweights} to $C$ defines weights
$\mathcal{D}\subseteq\R^e$: we call these the \textit{exceptional weights} of
$\Delta_g$. This terminology is due to the following result.
\begin{theorem}\label{thm:laplaceresults}
Let $(L,g)$ be a conifold with $e$ ends. Let $\mathcal{D}$ denote
the exceptional weights of $\Delta_g$. Then $\mathcal{D}$ is a discrete subset of $\R^e$ and the Laplace operator
\begin{equation*}
\Delta_g:W^p_{k,\boldsymbol{\beta}}(L)\rightarrow
W^p_{k-2,\boldsymbol{\beta}-2}(L)
\end{equation*}
is Fredholm iff $\boldsymbol{\beta}\notin \mathcal{D}$.
\end{theorem}
The above theorem, coupled with the ``change of index formula", leads to the
following conclusion, cf. \cite{pacini:weighted}.
\begin{corollary}\label{cor:laplaceresults}
Let $(L,g)$ be a compact Riemannian manifold. Consider the map
$\Delta_g:W^p_k(L)\rightarrow W^p_{k-2}(L)$. Then
\begin{equation*}
\mbox{Im}(\Delta_g)=\{u\in W^p_{k-2}(L):\int_L u\,\mbox{vol}_g=0\}, \ \
\mbox{Ker}(\Delta_g)=\R.
\end{equation*}
Let $(L,g)$ be an AC manifold. Consider the map
$\Delta_g:W^p_{k,\boldsymbol{\lambda}}(L)\rightarrow
W^p_{k-2,\boldsymbol{\lambda}-2}(L)$.
If $\boldsymbol{\lambda}>2-m$ is non-exceptional then this map is surjective. If
$\boldsymbol{\lambda}<0$ then this map is injective, so for
$\boldsymbol{\lambda}\in (2-m,0)$ it is an isomorphism.
Let $(L,g)$ be a CS manifold with $e$ ends. Consider the map
$\Delta_g:W^p_{k,\boldsymbol{\mu}}(L)\rightarrow
W^p_{k-2,\boldsymbol{\mu}-2}(L)$. If $\boldsymbol{\mu}\in (2-m,0)$ then
\begin{equation*}
\mbox{Im}(\Delta_g)=\{u\in W^p_{k-2,\boldsymbol{\mu}-2}(L): \int_L u\,vol_g=0\},
\ \ Ker(\Delta_g)=\R.
\end{equation*}
If $\boldsymbol{\mu}>0$ is non-exceptional then this map is injective and
\begin{equation*}
\mbox{dim(Coker($\Delta_g$))}=e+\sum_{0<\boldsymbol{\gamma}<\boldsymbol{\mu}}
m(\boldsymbol{\gamma}),
\end{equation*}
where $m(\boldsymbol{\gamma})$ is as in Definition \ref{def:exceptionalweights}.
Let $(L,g)$ be a CS/AC manifold with $s$ CS ends and $l$ AC ends. Consider the
map
\begin{equation*}
\Delta_g:W^p_{k,(\boldsymbol{\mu},\boldsymbol{\lambda})}(L)\rightarrow
W^p_{k-2,(\boldsymbol{\mu}-2,\boldsymbol{\lambda}-2)}(L).
\end{equation*}
If $(\boldsymbol{\mu},\boldsymbol{\lambda})\in (2-m,0)$ then this map is an
isomorphism. If $\boldsymbol{\mu}>0$ and $\boldsymbol{\lambda}<0$ are
non-exceptional then this map is injective and
\begin{equation*}
\mbox{dim(Coker($\Delta_g$))}=s+\sum_{0<\boldsymbol{\gamma}<\boldsymbol{\mu}}
m(\boldsymbol{\gamma}),
\end{equation*}
where $m(\boldsymbol{\gamma})$ is as in Definition \ref{def:exceptionalweights}.
Notice in particular that this dimension depends only on the harmonic functions
on the CS cones.
\end{corollary}
\section{Moduli spaces of special Lagrangian conifolds}\label{s:moduli}
Recall the statement of the Implicit Function Theorem.
\begin{theorem} \label{thm:IFT}
Let $F:E_1\rightarrow E_2$ be a smooth map between Banach spaces such that
$F(0)=0$. Assume $P:=dF[0]$ is surjective and $\mbox{Ker}(P)$ admits a closed
complement $Z$, \textit{i.e.} $E_1=\mbox{Ker}(P)\oplus Z$. Then there exists a
smooth map $\Phi:\mbox{Ker}(P)\rightarrow Z$ such that $F^{-1}(0)$ coincides
locally with the graph $\Gamma(\Phi)$ of $\Phi$. In particular, $F^{-1}(0)$ is
(locally) a smooth Banach submanifold of $E_1$.
\end{theorem}
The following result is straight-forward.
\begin{prop}\label{prop:fredholmreducestofinitedim}
Let $F:E_1\rightarrow E_2$ be a smooth map between Banach spaces such that
$F(0)=0$. Assume $P:=dF[0]$ is Fredholm. Set $\mathcal{I}:=\mbox{Ker}(P)$ and
choose $Z$ such that $E_1=\mathcal{I}\oplus Z$. Let $\mathcal{O}$ denote a
finite-dimensional subspace of $E_2$ such that $E_2=\mathcal{O}\oplus
\mbox{Im}(P)$. Define
\begin{equation*}
G:\mathcal{O}\oplus E_1\rightarrow E_2,\ \ (\gamma,e)\mapsto \gamma+F(e).
\end{equation*}
Identify $E_1$ with $(0,E_1)\subset \mathcal{O}\oplus E_1$. Then:
\begin{enumerate}
\item The map $dG[0]=Id\oplus P$ is surjective and
$\mbox{Ker}(dG[0])=\mbox{Ker}(P)$. Thus, by the Implicit Function Theorem, there
exist $\Phi:\mathcal{I}\rightarrow\mathcal{O}\oplus Z$ such that
$G^{-1}(0)=\Gamma(\Phi)$.
\item $F^{-1}(0)=\{(i,\Phi(i)):\Phi(i)\in
Z\}=\{(i,\Phi(i)):\pi_{\mathcal{O}}\circ\Phi(i)=0\}$,
where $\pi_{\mathcal{O}}:\mathcal{O}\oplus Z\rightarrow\mathcal{O}$ is the
standard projection.
\item Let $\pi_{\mathcal{I}}:\mathcal{I}\oplus Z\rightarrow \mathcal{I}$ denote
the standard projection. Then $\pi_{\mathcal{I}}$ is a continuous open map so it
restricts to a homeomorphism
\begin{equation*}
\pi_{\mathcal{I}}:F^{-1}(0)\rightarrow (\pi_{\mathcal{O}}\circ\Phi)^{-1}(0)
\end{equation*}
between $F^{-1}(0)$ and the zero set of the smooth map
$\pi_{\mathcal{O}}\circ\Phi:\mathcal{I}\rightarrow\mathcal{O}$, which is defined
between finite-dimensional spaces.
\end{enumerate}
\end{prop}
We now have all the ingredients necessary to prove various smoothness results
for SL moduli spaces. In all cases we follow the same steps. Section
\ref{s:setup} described each moduli space as the zero set of a map $F$. The
first step is to use regularity to show that one can equivalently study the zero
set of a map $\tilde{F}$. The domain of $\tilde{F}$ is of the form $K\times
W^p_{k,(\boldsymbol{\mu},\boldsymbol{\lambda})}(L)$ where $K$ is a
finite-dimensional vector space defined in terms of spaces introduced in
Sections \ref{ss:closedforms} and \ref{s:setup}. Roughly speaking, this
corresponds to separating the obvious Hamiltonian deformations of $L$ from a
finite-dimensional space of other Lagrangian deformations. The geometric
description of the latter depends on the case in question. The differential
$d\tilde{F}[0]$ is then a finite-dimensional perturbation of the Laplace
operator $\Delta_g$ acting on functions. The second step is to analyze this
linearized operator, showing that under appropriate conditions it is
surjective. The third step is to identify the kernel of $d\tilde{F}[0]$, at
least up to projections. One can then apply the Implicit Function Theorem and
conclude.
\subsubsection*{Smooth compact special Lagrangians} The following result was first proved by McLean \cite{mclean}.
\begin{theorem} \label{thm:mclean}
Let $L$ be a smooth compact SL submanifold of a CY manifold $M$. Let
$\mathcal{M}_L$ denote the moduli space of SL deformations of $L$. Then
$\mathcal{M}_L$ is a smooth manifold of dimension $b^1(L)$.
\end{theorem}
\begin{proof}
Choose $k\geq 3$ and $p>m$. Consider the space $\mbox{Ker}(d)$ of closed 1-forms
in $W^p_{k-1}(\Lambda^1)$. Let $\mathcal{D}_L$ denote the forms
$\alpha\in\mbox{Ker}(d)$ whose graph $\Gamma(\alpha)$ lies in $\mathcal{U}$.
Notice that $\Gamma(\alpha)$ is a well-defined $C^1$ Lagrangian submanifold in
$\mathcal{U}$ by the standard Sobolev embedding
$W^p_{k-1}(\Lambda^1)\hookrightarrow C^1(\Lambda^1)$. For the same reason,
$\mathcal{D}_L$ is an open neighbourhood of the origin in $\mbox{Ker}(d)$.
Consider the map
\begin{equation}\label{eq:cptFsobolev}
F:\mathcal{D}_L\rightarrow \{u\in W^p_{k-2}(L):\int_L u\,vol_g=0\}, \ \
\alpha\mapsto \star(\pi_*((\Phi_L^*\Imag\,\Omega)_{|\Gamma(\alpha)})).
\end{equation}
Recall that $W^p_{k-2}(L)$ is closed under multiplication.
Together with the ideas of Proposition \ref{prop:cptnonlinear}, this shows that
$F$ is a (locally well-defined) smooth map between Banach spaces with
differential $dF[0](\alpha)=d^*\alpha$. Assume $\alpha\in F^{-1}(0)$. Then, by
composition with $\Phi_L$, $\alpha$ defines a $C^1$ SL submanifold in $M$.
Standard regularity results for minimal submanifolds then show that $\alpha$ is
smooth. Thus $\mathcal{M}_L$ is locally homeomorphic, via $\Phi_L$, to
$F^{-1}(0)$.
Decomposition \ref{decomp:closedforms} shows that any $\alpha\in F^{-1}(0)$ is
of the form $\alpha=\beta+df$ for some unique $\beta\in H$ and some $f\in
C^\infty(L)$, defined up to a constant. We can thus re-phrase the SL deformation
problem as follows. Define $\mathcal{\tD}_L$ as the space of pairs $(\beta,f)$
in $H\times W^p_k(L)$ such that $\alpha:=\beta+df\in \mathcal{D}_L$. Clearly
$\mathcal{\tD}_L$ is an open neighbourhood of the origin. Then $\mathcal{\tD}_L$
is the domain of the (locally defined) map between Banach spaces
\begin{equation}\label{eq:cptFsobolevbis}
\tilde{F}:H\times W^p_k(L)\rightarrow \{u\in W^p_{k-2}(L):\int_L u\,vol_g=0\},\
\ \tilde{F}(\beta,f):=F(\beta+df).
\end{equation}
Clearly, $d\tilde{F}[0](\beta,f)=d^*\beta+\Delta_g f$. Let $\R$ denote the space
of constant functions in $W^p_k(L)$. Notice that both $\mathcal{\tD}_L$ and
$\tilde{F}$ are invariant under translations in $\R$. Assume
$\tilde{F}(\beta,f)=0$. With respect to $f$ this is a second-order elliptic
equation. Standard regularity results show that $f$ is smooth. This proves that
$\mathcal{M}_L$ is locally homeomorphic to the quotient space
$\tilde{F}^{-1}(0)/\R$. To conclude, it is sufficient to prove that
$\tilde{F}^{-1}(0)$ is smooth. According to Corollary \ref{cor:laplaceresults},
the map
\begin{equation}
\Delta_g:W^p_k(L)\rightarrow \{u\in W^p_{k-2}(L):\int_Lu\,vol_g=0\}
\end{equation}
is surjective. This implies that $d\tilde{F}[0]$ is surjective. Let $\beta_i$ be
a basis for $H$. For each $\beta_i$ the equation $d\tilde{F}[0](\beta_i,f)=0$
admits a solution $f_i$. Another solution is given by the pair $\beta=0$, $f=1$.
It is simple to check that these give a basis for the kernel of $d\tilde{F}[0]$.
Applying the Implicit Function Theorem we conclude that $\tilde{F}^{-1}(0)$ is
smooth of dimension $b^1(L)+1$, thus $\mathcal{M}_L$ is smooth of dimension
$b^1(L)$.
\end{proof}
\subsubsection*{AC special Lagrangians} The analogous result for AC SLs was originally proved independently by the
author \cite{pacini:defs} and by Marshall \cite{marshall}. We present here a
simplified proof, starting with the following weighted regularity result due to
Joyce, cf. \cite{joyce:I} Theorems 5.1 and 7.7.
\begin{lemma}\label{l:SLweightedregularity}
Let $\mathcal{C}$ be a SL cone in $\C^m$, endowed with the induced metric
$\tilde{g}$ and orientation. Define $\Phi_{\mathcal{C}}:\mathcal{U}\rightarrow
\C^m$ and the map $F$ as in Section \ref{ss:coneslsetup}. Fix any $\mu>2$ and
$\lambda<2$ with $\lambda\neq 0$. Assume given a closed 1-form $\alpha\in
C^1_{(\mu-1,\lambda-1)}(\mathcal{U})$ satisfying $F(\alpha)=0$. Analogously to
Decomposition \ref{decomp:closedforms_accs}, we can write $\alpha=\alpha'+dA$
where (i) $\alpha'$ is compactly-supported on the small end and
translation-invariant on the large end, and (ii) $A\in C^1_{(\mu,\lambda)}(L)$.
Then $\alpha'$ is smooth and $A\in C^\infty_{(\mu,\lambda)}(L)$, so $\alpha\in
C^\infty_{(\mu-1,\lambda-1)}(\mathcal{U})$.
\end{lemma}
\begin{proof} Standard regularity results for minimal submanifolds show that
$\alpha\in C^1_{(\mu-1,\lambda-1)}(\mathcal{U})\cap C^\infty(\mathcal{U})$.
Using the same ideas as in the proof of Decomposition
\ref{decomp:closedforms_accs}, this suffices to prove that $\alpha'$ and $A$ are
smooth. It is thus enough to show that the higher derivatives of $A$ converge at
the correct rate as $r\rightarrow\infty$ and $r\rightarrow 0$. We sketch here a
proof for $r\rightarrow\infty$, referring to \cite{joyce:I} for details; the
other case is analogous.
In terms of $A$, \textit{i.e.} absorbing the $\alpha'$-terms into the operator,
the equation $F(\alpha)=0$ corresponds to an equation $\tilde{F}(A)=0$. Given
$r_0>0$ and $\epsilon<<1$, consider the equivalent equation
\begin{equation}\label{eq:conformalSLeq}
r^2\tilde{F}(A)=0\ \ \mbox{restricted to }\Sigma\times(r_0-\epsilon
r_0,r_0+\epsilon r_0).
\end{equation}
As in Theorem \ref{thm:mclean} the linearization of $\tilde{F}$ is
$\Delta_{\tilde{g}}$. One can check that the change of coordinates $r=e^z$
transforms Equation \ref{eq:conformalSLeq} into an equation of the form
\begin{equation}\label{eq:conformalSLeqbis}
\Delta_{\tilde{h}}(A)+\dots=0\ \ \mbox{restricted to
}\Sigma\times(r_0'-\epsilon', r_0'+\epsilon'),
\end{equation}
where $\tilde{h}$ is the ``cylindrical metric"
$\tilde{h}:=r^{-2}\tilde{g}=dz^2+g'$. Up to a translation we can identify
$\Sigma\times(r_0'-\epsilon', r_0'+\epsilon')$ with the fixed, \textit{i.e.}
$r_0$-independent, domain $\Sigma\times(-\epsilon', \epsilon')$. One can show
that Equation \ref{eq:conformalSLeqbis} converges to the equation
$\Delta_{\tilde{h}}(A)=0$ on this domain in such a way that interior estimates
for the solutions are uniform as $r_0\rightarrow\infty$. In particular, in terms
of H\"{o}lder norms, there exists a constant $C=C(k,\beta)$ independent of $r_0$
such that
\begin{equation}\label{eq:SLregestimate}
\|A\|_{C^{k,\beta}}\leq C\cdot\|A\|_{C^1}
\end{equation}
on the domain $\Sigma\times(-\epsilon', \epsilon')$ and with respect to the
metric $\tilde{h}$. To be precise, as this is an ``interior'' estimate, the
domain on the left hand side is slightly smaller than the domain on the right
hand side.
Let us now write this estimate in terms of the coordinate $r$ and multiply both
sides by $r^{-\lambda}$. We can then check that
\begin{equation}\label{eq:SLregestimatebis}
\|A\|_{C^k_\lambda}\leq C\cdot\|A\|_{C^1_\lambda}
\end{equation}
on the domain $\Sigma\times(r_0-\epsilon r_0, r_0+\epsilon r_0)$ and with
respect to the metric $\tilde{g}$. As $r_0$ is arbitrary and
$\|A\|_{C^1_\lambda}$ is bounded on the large end, this shows that
$\|A\|_{C^k_\lambda}$ is bounded for all $k$ so $A\in C^\infty_\lambda$.
\end{proof}
\begin{theorem} \label{thm:pacini}
Let $L$ be an AC SL submanifold of $\C^m$ with rate
$\boldsymbol{\lambda}$. Let $\mathcal{M}_L$ denote the moduli space of SL
deformations of $L$ with rate $\boldsymbol{\lambda}$. Consider the operator
\begin{equation}\label{eq:aclaplacian}
\Delta_g:W^p_{k,\boldsymbol{\lambda}}(L)\rightarrow
W^p_{k-2,\boldsymbol{\lambda}-2}(L).
\end{equation}
\begin{enumerate}
\item If $\boldsymbol{\lambda}\in (0,2)$ is a non-exceptional weight for
$\Delta_g$ then $\mathcal{M}_L$ is a smooth manifold of dimension
$b^1(L)+\mbox{dim(Ker$(\Delta_g)$)}-1$.
\item If $\boldsymbol{\lambda}\in (2-m,0)$ then $\mathcal{M}_L$ is a smooth
manifold of dimension $b^1_c(L)$.
\end{enumerate}
\end{theorem}
\begin{proof}
As in Theorem \ref{thm:mclean}, choose $k\geq 3$ and $p>m$ so that
$W^p_{k-1,\boldsymbol{\lambda}-1}(\Lambda^1)\subset
C^1_{\boldsymbol{\lambda}-1}(\Lambda^1)$. Let $\mathcal{D}_L$ denote the space
of closed 1-forms in $W^p_{k-1,{\boldsymbol{\lambda}-1}}(\Lambda^1)$ whose graph
$\Gamma(\alpha)$ lies in $\mathcal{U}$. Consider the map
\begin{equation}\label{eq:acFsobolev}
F:\mathcal{D}_L\rightarrow W^p_{k-2,\boldsymbol{\lambda}-2}(L), \ \
\alpha\mapsto \star(\pi_*((\Phi_L^*\Imag\,\tilde{\Omega})_{|\Gamma(\alpha)})).
\end{equation}
Assume $\boldsymbol{\lambda}<2$. In this case Theorem
\ref{thm:embedding} shows that
$W^p_{k-2,\boldsymbol{\lambda}-2}(L)$ is closed under
multiplication. Together with the ideas of Proposition
\ref{prop:cptnonlinear}, this shows that $F$ is a (locally well-defined) smooth
map between Banach spaces with differential $dF[0](\alpha)=d^*\alpha$. Assume
$F(\alpha)=0$. Theorem \ref{thm:embedding} and Lemma
\ref{l:SLweightedregularity} then show that $\alpha\in
C^\infty_{\boldsymbol{\lambda}-1}(\Lambda^1)$ so $F^{-1}(0)$ is locally
homeomorphic, via $\Phi_L$, to $\mathcal{M}_L$.
Now assume $\boldsymbol{\lambda}\in (0,2)$. Decomposition
\ref{decomp:closedforms_growth} shows that any $\alpha\in F^{-1}(0)$ is of the
form $\alpha=\beta+df$, for some $\beta\in H$ and some $f\in
C^\infty_{\boldsymbol{\lambda}}(L)$. Define $\mathcal{\tD}_L$ as the space of
pairs $(\beta,f)$ in $H\times W^p_{k,\boldsymbol{\lambda}}(L)$ such that
$\alpha:=\beta+df\in \mathcal{D}_L$. Clearly $\mathcal{\tD}_L$ is an open
neighbourhood of the origin. Then $\mathcal{\tD}_L$ is the domain of the
(locally defined) smooth map between Banach spaces
\begin{equation}\label{eq:acFsobolevbis}
\tilde{F}:H\times W^p_{k,\boldsymbol{\lambda}}(L)\rightarrow
W^p_{k-2,\boldsymbol{\lambda}-2}(L),\ \ \tilde{F}(\beta,f):=F(\beta+df)
\end{equation}
with $d\tilde{F}[0](\beta,f)=d^*\beta+\Delta_g f$ and invariant under
translations in $\R$. Assume $\tilde{F}(\beta,f)=0$. Theorem \ref{thm:embedding}
and Lemma \ref{l:SLweightedregularity} then show that $f\in
C^\infty_{\boldsymbol{\lambda}}(L)$. This proves that $\mathcal{M}_L$ is locally
homeomorphic, via $\Phi_L$, to the quotient space $\tilde{F}^{-1}(0)/\R$. To
conclude, it is thus sufficient to prove that $\tilde{F}^{-1}(0)$ is smooth. For
this we need to further assume that $\boldsymbol{\lambda}$ is non-exceptional.
Then Corollary \ref{cor:laplaceresults} shows that the map of Equation
\ref{eq:aclaplacian} is surjective, so $d\tilde{F}[0]$ is surjective. Let
$\beta_i$ be a basis for $H$. For each $\beta_i$ the equation
$d\tilde{F}[0](\beta_i,f)=0$ admits a solution $f_i$. More solutions are given
by the pairs $\beta=0$, $f\in Ker(\Delta_g)$. It is simple to check that these
give a basis for the kernel of $d\tilde{F}[0]$. Applying the Implicit Function
Theorem we conclude that $\tilde{F}^{-1}(0)$ is smooth of dimension
$\mbox{dim}(H\oplus\mbox{Ker}(\Delta_g))$. Thus $\mathcal{M}_L$ is smooth and
has the claimed dimension.
Now assume $\boldsymbol{\lambda}\in (2-m,0)$. In this case Decomposition
\ref{decomp:closedforms_decay} shows that any $\alpha\in F^{-1}(0)$ is of the
form $\alpha=\beta+dv+df$, for some $\beta\in \widetilde{H}_\infty$, $dv\in
d(E_\infty)$ and $df\in d(C^\infty_{\boldsymbol{\lambda}}(L))$. We can use
regularity as before to prove that $\mathcal{M}_L$ is locally homeomorphic to
the quotient space $\tilde{F}^{-1}(0)/\R$, for the (locally defined) map
\begin{equation}\label{eq:acFsobolevter}
\tilde{F}:\widetilde{H}_\infty\times E_\infty\times
W^p_{k,\boldsymbol{\lambda}}(L)\rightarrow W^p_{k-2,\boldsymbol{\lambda}-2}(L),\
\ \tilde{F}(\beta,v,f)=F(\beta+dv+df).
\end{equation}
Notice that this time the constant functions $\R$ are contained in $E_\infty$.
We conclude as before that $\tilde{F}^{-1}(0)$ is smooth, this time of dimension
$\mbox{dim}(\widetilde{H}_\infty\oplus E_\infty)$. Remark
\ref{rem:compactsupport} then shows that $\mathcal{M}_L$ is smooth of dimension
$b^1_c(L)$.
\end{proof}
\subsubsection*{CS special Lagrangians} Now assume that $L$ is CS SL with singularities modelled on cones
$\mathcal{C}_i$. It turns out that smoothness of $\mathcal{M}_L$ then requires
an additional ``stability" assumption on $\mathcal{C}_i$. Roughly speaking, it
is required that the cones $\mathcal{C}_i$ admit no additional harmonic
functions with prescribed growth, beyond those which necessarily exist for
geometric reasons.
\begin{definition}\label{def:stability}
Let $\mathcal{C}$ be a SL cone in $\C^m$. Let $(\Sigma,g')$ denote the link of
$\mathcal{C}$ with the induced metric. Assume $\mathcal{C}$ has a unique
singularity at the origin; equivalently, assume that $\Sigma$ is smooth and that
it is not a sphere $\Sph^{m-1}\subset \Sph^{2m-1}$. Recall from the proof of
Proposition \ref{prop:csnonlinear} that the standard action of $\sunitary
m\ltimes\C^m$ on $\C^m$ admits a moment map $\mu$ and that the components of
$\mu$ restrict to harmonic functions on $\mathcal{C}$. Let $G$ denote the
subgroup of $\sunitary m$ which preserves $\mathcal{C}$. Then $\mu$ defines on
$\mathcal{C}$ $2m$ linearly independent harmonic functions of linear growth; in
the notation of Definition \ref{def:exceptionalweights} these functions are
contained in the space $V_\gamma$ with $\gamma=1$. The moment map also defines
on $\mathcal{C}$ $m^2-1-\mbox{dim}(G)$ linearly independent harmonic functions
of quadratic growth: these belong to the space $V_\gamma$ with $\gamma=2$.
Constant functions define a third space of homogeneous harmonic functions on
$\mathcal{C}$, \textit{i.e.} elements in $V_\gamma$ with $\gamma=0$. In
particular, these three values of $\gamma$ are always exceptional values for the
operator $\Delta_{\tilde{g}}$ on any SL cone, in the sense of Definition
\ref{def:exceptionalweights}.
We say that $\mathcal{C}$ is \textit{stable} if these are the only functions in
$V_\gamma$ for $\gamma=0,1,2$ and if there are no other exceptional values
$\gamma$ in the interval $[0,2]$. More generally, let $L$ be a CS or CS/AC SL
submanifold. We say that a singularity $x_i$ of $L$ is \textit{stable} if the
corresponding cone $\mathcal{C}_i$ is stable.
\end{definition}
The following result is due to Joyce \cite{joyce:II}.
\begin{theorem} \label{thm:joyce}
Let $L$ be a CS SL submanifold of $M$ with $s$ singularities and rate
$\boldsymbol{\mu}$. Let $\mathcal{M}_L$ denote the moduli space of SL
deformations of $L$ with moving singularities and rate $\boldsymbol{\mu}$.
Assume $\boldsymbol{\mu}$ is non-exceptional for the map
\begin{equation}\label{eq:cslaplacian}
\Delta_g:W^p_{k,\boldsymbol{\mu}}(L)\rightarrow \{u\in
W^p_{k-2,\boldsymbol{\mu}-2}(L): \int_L u\,vol_g=0\}.
\end{equation}
Then $\mathcal{M}_L$ is locally homeomorphic to the zero set of a smooth map
$\Phi: \mathcal{I}\rightarrow\mathcal{O}$
defined (locally) between finite-dimensional vector spaces. If
$\boldsymbol{\mu}=2+\epsilon$ and all singularities are stable then
$\mathcal{O}=\{0\}$ and $\mathcal{M}_L$ is smooth of dimension
$\mbox{dim}(\mathcal{I})=b^1_c(L)-s+1$.
\end{theorem}
\begin{proof}
Start with a map $F$ defined as in Section \ref{ss:accsslsetup} on
$\tilde{\E}\times W^p_{k-1,\boldsymbol{\mu}-1}(\Lambda^1)$. As in Theorem
\ref{thm:pacini}, regularity and Decomposition \ref{decomp:closedforms_decay}
show that $\mathcal{M}_L$ is locally homeomorphic to $\tilde{F}^{-1}(0)/\R$,
where $\tilde{F}$ is the (locally-defined) map
\begin{eqnarray*}
\tilde{F}:\tilde{\E}\times \widetilde{H}_0\times E_0\times
W^p_{k,\boldsymbol{\mu}}(L) &\rightarrow& \{u\in
W^p_{k-2,\boldsymbol{\mu}-2}(L):\int_Lu\,vol_g=0\}\\
(\tilde{e},\beta,v,f) &\mapsto& F(\tilde{e},\beta+dv+df),
\end{eqnarray*}
invariant under translations in $\R\subset E_0$. As in Proposition
\ref{prop:csnonlinear},
$d\tilde{F}[0](y,\beta,v,f)=d^*\beta+\Delta_g(\chi(y)+v+f)$. Now consider the
restricted map
\begin{equation}\label{eq:csrestrictedlin}
d\tilde{F}[0]:T_e\tilde{\E}\oplus E_0\oplus
W^p_{k,\boldsymbol{\mu}}(L)\rightarrow \{u\in
W^p_{k-2,\boldsymbol{\mu}-2}(L):\int_Lu\,vol_g=0\}.
\end{equation}
We claim that the kernel of this map is given by the constant functions $\R$. To
prove this, assume $d\tilde{F}[0](\chi(y)+v+f)=0$. Since $\chi(y)+v+f\in
W^p_{k,-\boldsymbol{\epsilon}}(L)$, Corollary \ref{cor:laplaceresults} shows
that $\chi(y)+v+f$ is constant, \textit{i.e.} $d(\chi(y)+v+f)=0$. In other words
the infinitesimal Lagrangian deformation of $L$ defined by $(y,v,f)$ is trivial,
so in particular $y=0$. This implies $\chi(y)=0$ and it is simple to conclude
that $f=0$ and $v\in\R$.
Let $\mathcal{O}$ denote the cokernel of the map of Equation
\ref{eq:csrestrictedlin}. More precisely, we define it to be a
finite-dimensional space of $W^p_{k-2,\mu-2}(L)$ such that
\begin{equation}
\mathcal{O}\oplus d\tilde{F}[0]\left(T_e\tilde{\E}\oplus E_0\oplus
W^p_{k,\boldsymbol{\mu}}\right)=\{u\in
W^p_{k-2,\boldsymbol{\mu}-2}(L):\int_Lu\,vol_g=0\}.
\end{equation}
Consider the map
\begin{eqnarray*}
G:\mathcal{O}\times \tilde{\E}\times \widetilde{H}_0\times E_0\times
W^p_{k,\boldsymbol{\mu}}(L) &\rightarrow& \{u\in
W^p_{k-2,\boldsymbol{\mu}-2}(L):\int_Lu \,vol_g=0\}\\
(\gamma,\tilde{e},\beta,v,f) &\mapsto& \gamma+\tilde{F}(\tilde{e},\beta,v,f).
\end{eqnarray*}
Again, $G$ is invariant under translations in $\R$. By construction, the
restriction of $dG[0]$ to the space $\mathcal{O}\oplus T_e\tilde{\E}\oplus
E_0\oplus W^p_{k,\boldsymbol{\mu}}$ is surjective with kernel $\R$. We now have
the following information about the map $G$. Firstly,
$\mbox{Ker}(dG[0])=V\oplus\R$, where $V$ is some vector space projecting
isomorphically onto $\widetilde{H}_0$. Secondly, by the Implicit Function
Theorem, the set $G^{-1}(0)$ is smooth and can be locally written as the graph
of a smooth map $\Phi$ defined on the kernel of $dG[0]$, thus on
$\widetilde{H}_0\oplus\R$.
As in Proposition \ref{prop:fredholmreducestofinitedim} we can conclude that the
projection onto $\widetilde{H}_0\oplus\R$ restricts to a homeomorphism
$\tilde{F}^{-1}(0)\simeq(\pi_{\mathcal{O}}\circ\Phi)^{-1}(0)$. It is simple to
check that $\Phi$ is invariant under translations in $\R$.
Restricting $\Phi$ to $\mathcal{I}:=\widetilde{H}_0$ proves the first
claim.
Now let us further assume that $\boldsymbol{\mu}=2+\epsilon$ and that all
singularities are stable. Here, $\epsilon$ is to be understood as in Remark
\ref{rem:muregularity}; in particular, the moduli space we will obtain is
independent of the particular $\epsilon$ chosen. Recall from Corollary
\ref{cor:laplaceresults} that for $\boldsymbol{\mu}>2-m$ we can compute the
dimension of $\mbox{Coker}(\Delta_g)$ in terms of the number of harmonic
functions on the cones $\mathcal{C}_i$. Recall from Definition
\ref{def:stability} that SL cones always admit a certain number of harmonic
functions. This implies that, for the operator
$\Delta_g:W^p_{k,\boldsymbol{\mu}}(L)\rightarrow
W^p_{k-2,\boldsymbol{\mu}-2}(L)$,
\begin{equation}
\mbox{dim(Coker$(\Delta_g)$)}\geq d, \ \ \mbox{where } d:=\sum_{i=1}^e
\left(1+2m+m^2-1-\mbox{dim}(G_i)\right).
\end{equation}
The stability condition is equivalent to $\mbox{dim(Coker$(\Delta_g)$)}=d$. This
means that the cokernel of the operator in Equation \ref{eq:cslaplacian} has
dimension $d-1$. Notice that $d$ is also the dimension of the space
$T_e\tilde{\E}\oplus E_0$. Our calculation of the kernel thus implies that the
map $d\tilde{F}[0]$ of Equation \ref{eq:csrestrictedlin} is surjective. Thus
$\mathcal{O}=\{0\}$. We can now apply the Implicit Function Theorem directly to
$\tilde{F}$ to obtain that $\tilde{F}^{-1}(0)$ is smooth, of dimension
$\mbox{dim}(\widetilde{H}_0)+1$. Quotienting by $\R$ and using Equation
\ref{eq:dimexactseq} gives the desired result.
\end{proof}
We call $\mathcal{O}$ the \textit{obstruction space} of the SL deformation
problem.
\subsubsection*{CS/AC special Lagrangians in $\C^m$} We can now state and prove the main result of this paper.
\begin{theorem} \label{thm:accssl}
Let $L$ be a CS/AC SL submanifold of $\C^m$ with $s$ CS ends, $l$ AC ends
and rate $(\boldsymbol{\mu},\boldsymbol{\lambda})$. Let $\mathcal{M}_L$ denote
the moduli space of SL deformations of $L$ with moving singularities and rate
$(\boldsymbol{\mu},\boldsymbol{\lambda})$. Assume
$(\boldsymbol{\mu},\boldsymbol{\lambda})$ is non-exceptional for the map
\begin{equation}\label{eq:accslaplacian}
\Delta_g:W^p_{k,(\boldsymbol{\mu},\boldsymbol{\lambda})}(L)\rightarrow
W^p_{k-2,(\boldsymbol{\mu}-2,\boldsymbol{\lambda}-2)}(L).
\end{equation}
We will restrict our attention to the two cases $\boldsymbol{\lambda}\in
(2-m,0)$ or $\boldsymbol{\lambda}\in (0,2)$. In either case $\mathcal{M}_L$ is
locally homeomorphic to the zero set of a smooth map
$\Phi:\mathcal{I}\rightarrow\mathcal{O}$ defined (locally) between
finite-dimensional vector spaces. If furthermore $\boldsymbol{\mu}=2+\epsilon$
and all singularities are stable then $\mathcal{O}=\{0\}$ and $\mathcal{M}_L$ is
smooth of dimension $\mbox{dim}(\mathcal{I})$. Specifically:
\begin{enumerate}
\item If $\boldsymbol{\lambda}\in (2-m,0)$ then
$\mbox{dim}(\mathcal{I})=b^1_c(L)-s$.
\item If $\boldsymbol{\lambda}\in (0,2)$ then
$\mbox{dim}(\mathcal{I})=b^1_{c,\bullet}(L)-s+\sum_{i=1}^l d_i$,
where $d_i$ is the number of harmonic functions on the AC end $S_i$ of the form
$r^\gamma\sigma(\theta)$ with $\gamma\in [0,\lambda_i]$.
\end{enumerate}
\end{theorem}
\begin{proof}
Start with a map $F$ defined as in the previous theorems on $\tilde{\E}\times
W^p_{k-1,(\boldsymbol{\mu}-1,\boldsymbol{\lambda}-1)}(\Lambda^1)$, such that
$\mathcal{M}_L\simeq F^{-1}(0)$. Let
$\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}}$ denote the map of Equation
\ref{eq:accslaplacian}.
We split the proof into two parts, depending on the range of
$\boldsymbol{\lambda}$. To begin, assume $\boldsymbol{\lambda}\in (2-m,0)$. By
regularity and Decomposition \ref{decomp:closedforms_accs}, $\mathcal{M}_L$ is
locally homeomorphic to $\tilde{F}^{-1}(0)/\R$, where
$\tilde{F}$ is the (locally-defined) map
\begin{eqnarray*}
\tilde{F}:\tilde{\E}\times \widetilde{H}_{0,\infty}\times E_{0,\infty}\times
W^p_{k,(\boldsymbol{\mu},\boldsymbol{\lambda})}(L) &\rightarrow&
W^p_{k-2,(\boldsymbol{\mu}-2,\boldsymbol{\lambda}-2)}(L)\\
(\tilde{e},\beta,v,f) &\mapsto& F(\tilde{e},\beta+dv+df).
\end{eqnarray*}
As in Proposition \ref{prop:accsnonlinear},
$d\tilde{F}[0](y,\beta,v,f)=d^*\beta+\Delta_g(\chi(y)+v+f)$. Now consider the
restricted map
\begin{equation}\label{eq:accsrestrictedlin}
d\tilde{F}[0]:T_e\tilde{\E}\oplus E_0\oplus
W^p_{k,(\boldsymbol{\mu},\boldsymbol{\lambda})}(L)\rightarrow
W^p_{k-2,(\boldsymbol{\mu}-2,\boldsymbol{\lambda}-2)}(L),
\end{equation}
where $E_0$ is the subspace of functions in $E_{0,\infty}$ which vanish on the
AC ends.
Notice that $\chi(y)+v+f\in
W^p_{k,(-\boldsymbol{\epsilon},\boldsymbol{\lambda})}(L)$. As in Theorem
\ref{thm:joyce} we can use Corollary \ref{cor:laplaceresults} to prove that the
map of Equation \ref{eq:accsrestrictedlin} is injective.
Let $\mathcal{O}$ denote the cokernel of the map of Equation
\ref{eq:accsrestrictedlin}. More precisely, we define it to be a
finite-dimensional subspace of
$W^p_{k-2,(\boldsymbol{\mu}-2,\boldsymbol{\lambda}-2)}(L)$ such that
\begin{equation}
\mathcal{O}\oplus d\tilde{F}[0]\left(T_e\tilde{\E}\oplus E_0\oplus
W^p_{k,(\boldsymbol{\mu},\boldsymbol{\lambda})}\right)=W^p_{k-2,(\boldsymbol{\mu
}-2,\boldsymbol{\lambda}-2)}(L).
\end{equation}
Consider the map
\begin{eqnarray*}
G:\mathcal{O}\times \tilde{\E}\times \widetilde{H}_{0,\infty}\times
E_{0,\infty}\times W^p_{k,(\boldsymbol{\mu},\boldsymbol{\lambda})}(L)
&\rightarrow& W^p_{k-2,(\boldsymbol{\mu}-2,\boldsymbol{\lambda}-2)}(L)\\
(\gamma,\tilde{e},\beta,v,f) &\mapsto& \gamma+\tilde{F}(\tilde{e},\beta,v,f).
\end{eqnarray*}
By construction the restriction of $dG[0]$ to the space $\mathcal{O}\oplus
T_e\tilde{\E}\oplus E_0\oplus W^p_{k,(\boldsymbol{\mu},\boldsymbol{\lambda})}$
is an isomorphism. Let $E'$ denote a complement of $E_0\oplus\R$ in
$E_{0,\infty}$, \textit{i.e.} $E_{0,\infty}=E_0\oplus\R\oplus E'$. As in Theorem
\ref{thm:joyce}, $G^{-1}(0)$ is smooth and can be written as the graph of a
smooth map $\Phi$ defined on $\widetilde{H}_{0,\infty}\oplus(\R\oplus E')$.
Restricting $\Phi$ to $\mathcal{I}:=\widetilde{H}_{0,\infty}\oplus E'$ and using
the same arguments as in Proposition \ref{prop:fredholmreducestofinitedim} and
Theorem \ref{thm:joyce} then proves
the first claim regarding $\mathcal{M}_L$ for this range of
$\boldsymbol{\lambda}$. Notice that
$\mbox{dim}(\widetilde{H}_{0,\infty})=b^1_c(L)-(s+l)+1$ and $\mbox{dim}(E')=l-1$
so $\mbox{dim}(\mathcal{I})=b^1_c(L)-s$.
Now let us further assume that $\boldsymbol{\mu}=2+\epsilon$ and that all
singularities are stable. Here, as in Theorem \ref{thm:joyce}, $\epsilon$ is to
be understood as in Remark \ref{rem:muregularity}. By Corollary
\ref{cor:laplaceresults} and the definition of stability,
\begin{equation}
\mbox{dim(Coker$(\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}})$)}=d, \ \
\mbox{where } d:=\sum_{i=1}^s \left(1+2m+m^2-1-\mbox{dim}(G_i)\right).
\end{equation}
Again, $d$ is also the dimension of the space $T_e\tilde{\E}\oplus E_0$. Our
previous injectivity calculation thus implies that the map $d\tilde{F}[0]$ of
Equation \ref{eq:accsrestrictedlin} is an isomorphism. In particular,
$\mathcal{O}=\{0\}$. We can now apply the Implicit Function Theorem directly to
$\tilde{F}$ to obtain that $\tilde{F}^{-1}(0)$ is smooth. Quotienting by $\R$
shows that $\mathcal{M}_L$ is smooth.
We now start over again, under the assumption $\boldsymbol{\lambda}\in (0,2)$.
In this case we use the map
\begin{eqnarray*}
\tilde{F}:\tilde{\E}\times \widetilde{H}_{0,\bullet}\times E_0\times
W^p_{k,(\boldsymbol{\mu},\boldsymbol{\lambda})}(L) &\rightarrow&
W^p_{k-2,(\boldsymbol{\mu}-2,\boldsymbol{\lambda}-2)}(L)\\
(\tilde{e},\beta,v,f) &\mapsto& F(\tilde{e},\beta+dv+df)
\end{eqnarray*}
and the restricted map
\begin{equation}\label{eq:accsrestrictedlinbis}
d\tilde{F}[0]:T_e\tilde{\E}\oplus E_0\oplus
W^p_{k,(\boldsymbol{\mu},\boldsymbol{\lambda})}(L)\rightarrow
W^p_{k-2,(\boldsymbol{\mu}-2,\boldsymbol{\lambda}-2)}(L).
\end{equation}
Recall the construction of $E_0$ in Decomposition \ref{decomp:closedforms_accs}:
it is clear that we may assume that $\chi(T_e\tilde{\E})$ and $E_0$ are linearly
independent in $W^p_{k,(-\boldsymbol{\epsilon},-\boldsymbol{\epsilon})}(L)$.
Corollary \ref{cor:laplaceresults} proves that $\Delta_g$ is injective on this
space. Define a decomposition
\begin{equation}
T_e\tilde{\E}\oplus E_0=Z'\oplus Z''
\end{equation}
by imposing $\Delta_g(Z')=\Delta_g(T_e\tilde{\E}\oplus E_0)\cap
\mbox{Im}(\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}})$ and choosing any
complement $Z''$. Then one can check that the kernel of the map of Equation
\ref{eq:accsrestrictedlinbis} is isomorphic to
$Z'\oplus\mbox{Ker}(\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}})$.
Choose $\mathcal{O}$ in
$W^p_{k-2,(\boldsymbol{\mu}-2,\boldsymbol{\lambda}-2)}(L)$ such that
\begin{equation}
\mathcal{O}\oplus d\tilde{F}[0]\left(T_e\tilde{\E}\oplus E_0\oplus
W^p_{k,(\boldsymbol{\mu},\boldsymbol{\lambda})}\right)=W^p_{k-2,(\boldsymbol{\mu
}-2,\boldsymbol{\lambda}-2)}(L).
\end{equation}
Consider the map
\begin{eqnarray*}
G:\mathcal{O}\times \tilde{\E}\times \widetilde{H}_{0,\bullet}\times E_0\times
W^p_{k,(\boldsymbol{\mu},\boldsymbol{\lambda})}(L) &\rightarrow&
W^p_{k-2,(\boldsymbol{\mu}-2,\boldsymbol{\lambda}-2)}(L)\\
(\gamma,\tilde{e},\beta,v,f) &\mapsto& \gamma+\tilde{F}(\tilde{e},\beta,v,f).
\end{eqnarray*}
The restriction of $dG[0]$ to the space $\mathcal{O}\oplus T_e\tilde{\E}\oplus
E_0\oplus W^p_{k,(\boldsymbol{\mu},\boldsymbol{\lambda})}$ is surjective. As
before, this implies that $G^{-1}(0)$ can be parametrised via a smooth map
$\Phi$ defined (locally) on the space $\widetilde{H}_{0,\bullet}\oplus Z'\oplus
\mbox{Ker}(\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}})$. As usual, these
maps are invariant under translations in $\R\subset
Z'\oplus\mbox{Ker}(\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}})$. Setting
$\mathcal{I}:=(\widetilde{H}_{0,\bullet}\oplus Z'\oplus
\mbox{Ker}(\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}}))/\R$ and considering
the natural map on this quotient then proves the first claim regarding
$\mathcal{M}_L$ for this range of $\boldsymbol{\lambda}$.
Now assume that $\boldsymbol{\mu}=2+\epsilon$ and that all singularities are
stable. Choose $\boldsymbol{\lambda}'\in (2-m,0)$. We can restrict the map of
Equation \ref{eq:accsrestrictedlinbis} to the map
\begin{equation}\label{eq:accsrestrictedlinter}
d\tilde{F}[0]:T_e\tilde{\E}\oplus E_0\oplus
W^p_{k,(\boldsymbol{\mu},\boldsymbol{\lambda}')}(L)\rightarrow
W^p_{k-2,(\boldsymbol{\mu}-2,\boldsymbol{\lambda}'-2)}(L).
\end{equation}
Exactly as for Equation \ref{eq:accsrestrictedlin}, it is simple to prove that
Equation \ref{eq:accsrestrictedlinter} defines an isomorphism and that
$\mbox{dim}(T_e\tilde{\E}\oplus
E_0)=\mbox{dim(Coker($\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}'}$))}$,
where
\begin{equation*}
\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}'}:=\Delta_g:
W^p_{k,(\boldsymbol{\mu},\boldsymbol{\lambda}')}(L)\rightarrow
W^p_{k-2,(\boldsymbol{\mu}-2,\boldsymbol{\lambda}'-2)}(L).
\end{equation*}
One can check that the dimension of
$\mbox{Coker}(\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}})$ decreases as
$\boldsymbol{\lambda}$ increases. We can actually assume, cf.
\cite{pacini:weighted}, that
$\mbox{Coker}(\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}})\subseteq\mbox{
Coker}(\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}'})$. This proves that the
map of Equation \ref{eq:accsrestrictedlinbis} is surjective, \textit{i.e.}
$\mathcal{O}=\{0\}$, so $\tilde{F}^{-1}(0)$ and $\mathcal{M}_L$ are smooth. To
compute the dimension of this moduli space notice that $Z''\simeq
\mbox{Coker}(\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}})$ so
\begin{eqnarray}
\mbox{dim(Ker$(d\tilde{F}[0])$)} &=&
\mbox{dim(Ker$(\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}})$)}
+\mbox{dim}(Z')\nonumber\\
&=&
\mbox{dim(Ker$(\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}})$)}+\mbox{
dim(Coker($\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}'}$))}-\mbox{
dim(Coker($\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}}$))}\nonumber\\
&=&
i(\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}})-i(\Delta_{\boldsymbol{\mu},
\boldsymbol{\lambda}'}),
\end{eqnarray}
where $i$ denotes the index of the Fredholm map. This implies that the kernel of
the full map $d\tilde{F}[0]$ has dimension
$\mbox{dim}(\widetilde{H}_{0,\bullet})+i(\Delta_{\boldsymbol{\mu},\boldsymbol{
\lambda}})-i(\Delta_{\boldsymbol{\mu},\boldsymbol{\lambda}'})$. The conclusion
follows from Equation \ref{eq:cohomsequence_accsbis} and the change of index
formula, cf. \cite{pacini:weighted}.
\end{proof}
\begin{remark}
Notice that, when $\boldsymbol{\lambda}<0$ and the stability condition is
verified, the dimension of the SL moduli spaces appearing in Theorems
\ref{thm:mclean}, \ref{thm:pacini}, \ref{thm:joyce} and \ref{thm:accssl} is
purely topological. The cases analyzed in the theorems correspond exactly to the
cases analyzed in Corollary \ref{cor:topsummary}, in the sense that the moduli
spaces should be thought of as being modelled on the cohomology spaces which
appear in Corollary \ref{cor:topsummary}.
It is interesting to notice how decay conditions on AC and CS ends are
incorporated differently into these cohomology spaces: decay conditions on AC
ends correspond to using compactly-supported forms while decay conditions on CS
ends correspond to the condition that a certain restriction map vanishes.
Allowing $\boldsymbol{\lambda}>0$ changes the topological data, again in
agreement with Corollary \ref{cor:topsummary}. It also introduces new SL
deformations which depend on analytic data.
\end{remark}
\begin{example}
Let $\mathcal{C}$ be a SL cone in $\C^m$. Assume $\mathcal{C}$ is stable and
that its link $\Sigma$ is connected so that $s=1$. Using Poincar\'{e} Duality
and the fact that $\mathcal{C}\simeq\Sigma\times (0,\infty)$ we see that
\begin{equation}
b^1_c(\mathcal{C})=b^{m-1}(\mathcal{C})=b^{m-1}(\Sigma)=1.
\end{equation}
Theorem \ref{thm:accssl} then shows that, for $\lambda\in (2-m,0)$,
$\mathcal{M}_{\mathcal{C}}$ has dimension 0, \textit{i.e.} $\mathcal{C}$ is
rigid within this class of deformations.
Notice also that restriction defines isomorphisms $H^i(\mathcal{C};\R)\simeq
H^i(\Sigma;\R)$ so the long exact sequence \ref{eq:cohomsequence_accsbis}, using
$\Sigma_0=\Sigma$, leads to $H^i_{c,\bullet}(\mathcal{C};\R)=0$.
Theorem \ref{thm:accssl} then shows that $\mathcal{M}_{\mathcal{C}}$ has
dimension 0 if $\lambda\in (0,1)$ and has dimension $2m$ if $\lambda\in (1,2)$.
In the latter case the SL deformations are simply the translations of
$\mathcal{C}$ in $\C^m$.
\end{example}
\
{\bf Acknowledgments.} I would like to thank D. Joyce for many useful
explanations on his work and for his help and advice on various parts of this
paper. I would also like to thank S. Karigiannis for useful conversations. This
work was carried out while I was a Marie Curie EIF Fellow at the University of
Oxford.
\bibliographystyle{amsplain}
|
\section{Introduction}
While collisionless $N$-body simulations produce dark matter halos that are triaxial or prolate, simulations which include the effects of baryons result in more spherical or axisymmetric halos\citep{dubins_94, kkzanm04,deb_etal_08}. It has been suggested that the change in shape could be the consequence of chaotic scattering of the box orbits that form the "back bone" of triaxial galaxies \citep{ger_bin_85}. Alternatively, halo shapes might change due the adiabatic response of orbits to the change in the central potential. In a recent paper \citet{valluri_etal_10} showed that by applying a spectral method to analyze large numbers of randomly selected orbits in $N$-body simulations it was possible to clearly distinguish between these two options. In this paper we present a brief summary of some of their main results.
\section{Frequency Analysis of $N$-body orbits}
Prolate and triaxial dark matter halos were formed from the merger of spherical NFW halos \citep{nfw} and baryonic components (representing a disk, an elliptical galaxy or a massive compact nucleus) were grown adiabatically with time. The evolution of the halo was followed using {\sc pkdgrav} an efficient, multi-stepping, parallel tree code \citep{stadel_phd}. After the baryonic component was grown to full strength it was artifically "evaporated" to allows us to test for the importance of chaos in the evolution of the system \citep[][hereafter D08 and references therein]{deb_etal_08}.
In each system several thousand orbits were selected and their trajectories evolved in a frozen potential in each phase of the evolution of each halo. The initial triaxial/prolate phase is referred to as {\it phase a}, once the baryonic component is grown to full strength the halo is in {\it phase b}, and after the baryonic component had "evaporated" and the system returned to equilibrium the halo is in {\it phase c}. Each orbit was analyzed using a code that decompose its phase space trajectory to obtain its three fundamental frequencies of oscillation \citep{laskar_90, valluri_merritt_98}.
The fundamental frequencies were then used to obtain a complete picture of the properties of individual orbits: to distinguish between regular and chaotic orbits, to classify regular orbits into major orbit families, to quantify the average shape of an orbit and relate its shape to the shape of the halo, and to identify the major resonant families of the halo which determine its structure \citep{valluri_etal_10}.
\section{Results}
The largest of the three fundamental frequencies in each of the three phases of the evolution of our models (referred to as $\omega_a$, $\omega_b$ and $\omega_c$ respectively), can be compared to distinguish between adiabatic and chaotic evolution. In the case of a primarily adiabatic response, particles deep in the potential (large $\omega_a$) are expected to experience a greater increase in frequency than particles further from the center. In Figure~\ref{fig:omega_abc} we compare the frequencies of orbits in an initially triaxial halo ($\omega_a$) with their frequency in the presence of a baryonic component ($\omega_b$). For the extended disk ({\it Left}) (as well as other extended baryonic components), $\omega_b$ increases monotonically with $\omega_a$ with fairly small scatter, indicating that the orbits in this potential responded adiabatically (the dashed line shows the 1:1 correlation between the two frequencies).
\begin{figure}
\centering
\includegraphics[width=.37\textwidth]{001mvalluri_fig1a.ps}
\hfill
\includegraphics[width=.37\textwidth]{001mvalluri_fig1b.ps}
\caption{For $\sim$5000 particles frequencies before ($\omega_a$) versus after ($\omega_b$) the growth of a baryonic component are shown. {\it Left:} effect of an extended baryonic component (disk with 3~kpc scale length). {\it Right:} effect of a hard spherical point masses with 0.1~kpc softening.
\label{fig:omega_abc}}
\end{figure}
However, a hard central point mass ({\it Right}) results in significant scattering in $\omega_b$ with small values of $\omega_a$ (i.e. weakly bound orbits) having some of the largest values
of $\omega_b$. $\omega_b$ sometimes decreased instead of increasing providing further evidence for chaotic scattering in this case.
The orbital frequencies were used to distinguish between regular and chaotic orbits and to classify the regular orbits into major families. We studied both triaxial and prolate halos and our orbital analysis showed that the initial triaxial halo was composed of 84-86\% box orbits, 11-12\% long-axis tube orbits, 2\% short axis tubes, and 1-2\% chaotic orbits. In contrast the prolate halo had 15\% box orbits, 78\% long-axis tubes, 7\% short-axis tubes and no chaotic orbits. To test the hypothesis that it is the box orbits that are most significantly scattered by a central point mass we define the fractional change $\Delta\omega_{ac}$ in the frequency of an orbit from its value in the initial halo ($\omega_a$) to its value after the baryonic component was evaporated ($\omega_c$) to measure the amount of chaotic scattering.
Figure~\ref{fig:Pl3_jtot_rperi} shows that orbits with smaller pericenter radii $r_{\rm peri}$ experience a significantly larger change in frequency $\Delta \omega_{ac}$ than orbits at large pericenter radii. This is true for both the triaxial halo ({\it Left }) which is dominated by centrophilic box orbits and the prolate halo ({\it Right}) which is dominated by long-axis tubes.
Thus chaotic scattering is equally strong for the centrophobic long-axis tube orbits and centrophilic box orbits contrary to the prevailing view.
\begin{figure*}
\centering
\includegraphics[width=0.4\textwidth]{001mvalluri_fig2a.ps}
\hfill
\includegraphics[width=0.408\textwidth]{001mvalluri_fig2b.ps}
\caption{The change in frequency of an orbit $\omega_{ac}$ versus pericentric radii $r_{\rm peri}$ for triaxial halo ({\it Left}) and prolate halo ({\it Right}).
\label{fig:Pl3_jtot_rperi}}
\end{figure*}
In a triaxial potential in which the major, intermediate and short axes are along the Cartesian directions $x, y, z$ respectively, the oscillation frequencies satisfy the condition $
|\omega_x| < |\omega_y| < |\omega_z|$ for any orbit with the same over-all shape as the density distribution. For such an orbit we can define a shape parameter $\chi_s \equiv {|{\omega_y/\omega_z}|} - {|{\omega_x/\omega_z}|}$
which is positive for orbits with
elongation along the major axis of the figure, with larger values of $\chi_s$ implying a
greater degree of elongation.
Since orbits closer to the central potential are more significantly affected by
the baryonic component, we expect them to
become rounder ($\chi_s \rightarrow 0$) than orbits further out. Figure~\ref{fig:shape_peri} shows how the orbital shape parameter $\chi_s$ for various orbital types varies as a function of the pericenter radius $r_{\rm peri}$ for a triaxial halo ({\it Left}) and after a disk was grown in this halo with symmetry axis parallel to the short axis ({\it Right}). In each plot the curves show the mean shape distribution of orbits of a given orbital type at each radius, with orbital types indicated in the line-legends. In the initial triaxial halo boxes, long-axis tubes and chaotic orbits are significantly elongated ($\chi_s \sim 0.35$) both at small and large radii.
After the growth of the short-axis disk the orbits at small radii become axisymmetric ($\chi_s \sim 0$) while orbits at large radii (especially boxes) remain quite elongated. This change in orbital shape at small radii was seen in all halos regardless of the radial scale length of the baryonic component.
\begin{figure*}
\centering
\includegraphics[width=0.33\textwidth,angle=-90]{001mvalluri_fig3.ps}
\caption{The mean orbital shape parameter $\chi_s$ for different orbital types (indicated by line legends) vs. pericentric radius $r_{\rm peri}$.}
\label{fig:shape_peri}
\end{figure*}
\section{Implications}
The analysis of fundamental frequencies of orbits in $N$-body halos is a powerful technique that allows us to identify the primary physical processes that cause halo shapes to change in response to the growth of a baryonic component. We confirmed the conclusion reached by D08 that chaos is not an
important driver of shape evolution but found that significant
chaotic scattering does occur when the baryonic component is in the
form of a hard central point mass (of scale length $\sim 0.1$~kpc).
Regardless of the original orbital composition
of the triaxial or prolate halo, and regardless of the shape and radial
scale length of the baryonic component, halos become more oblate because orbits closer to
the center of the potential become axisymmetric. Although the resultant halos are almost oblate, their orbit populations contain orbits (boxes and long-axis tubes) that are characteristic of a triaxial rather than axisymmetric potential.
\begin{theacknowledgments}
MV would like to thank the organizers of the conference at UCLAN and the University of Malta for organizing an excellent meeting.
\end{theacknowledgments}
\bibliographystyle{aipproc}
|
\section{Introduction}
Low energy ion sputtering is a promising candidate for the fabrication of self-ordered nanostructures on both crystalline and amorphous surfaces \cite{ChasonJAP}. It was observed experimentally that ion sputtering at normal incidence or at oblique incidence can generate correlated dot or ripple patterns on metals \cite{ChasonPRBCu}, semiconductors \cite{ChasonGe, ChasonNIMBSi, GagoAPL2001} and insulators \cite{UmbachPRL}. The formation mechanism has been described theoretically by the competition between morphology induced surface instability and surface relaxation processes \cite{BradleyHarper, CuernoBarabasiPRL, MakeevNIMB, Munoz-GarciaPRB2008}.
Recently, it was reported by Ozaydin \etal \cite{OzaydinAPLMo} that molybdenum incorporation triggers the formation of ordered dot arrays on Si surfaces under the bombardment with Ar$^+$ ions at normal incidence while no correlated structures are generated in the absence of the impurities. Furthermore, Hofs\"{a}ss \etal \cite{Hofsaess} found that co-deposition of Au, Ag and Pt surfactants generates novel patterns and nanostructures on Si surfaces during ion sputtering which are absent without co-deposition. Moreover, Sanchez-Garcia \etal \cite{SanchezNanotech} observed a transition from hole to dot patterns on Si by tuning the amount of metal incorporation during ion sputtering at normal incidence. Additionally, Shenoy \etal \cite{ShenoyPRL} proposed that preferential sputtering can generate modulation both in surface morphology and composition. These findings add a new aspect to the present understanding of the development of surface topography during ion sputtering and suggest more detailed studies which address the involved roughening and smoothing mechanisms. Especially quantitative data on the influence of the metal incorporation on the morphology evolution are still missing.
In this contribution, a comparative experiment is described, which proves the important role of metal incorporation in triggering dot pattern formation on Si(100) surfaces and allows us to determine the Fe concentration at which the morphology evolution changes from roughening to smoothing.. Furthermore, the morphology evolution in the absence of metal incorporation clearly indicates that the relaxation process includes both $k^2$- and $k^4$-dependent smoothing mechanisms, with $k$ being the surface wavenumber.
\section{Experiment}
\label{sec:experiment}
The experiments were performed in an ultra-high-vacuum chamber with a base pressure of $1\times10^{-8}$ mbar. For generation of a low-pressure plasma, an inductively coupled RF plasma (ICP) generator is attached using a planar coil behind a quartz window of 150 mm diameter, which is immersed into the chamber \cite{Chevolleau2000}. At an Ar pressure of 4$\times$10$^{-3}$ mbar, the absorbed RF power was 100 W at a frequency of 13.56 MHz. Ions were extracted from the plasma by applying a 1200 eV negative DC bias to a flat sample holder which is mounted opposite to the quartz window at a distance of 10 cm. The electric field geometry and the low pressure guarantee an almost ideal normal incidence of the ions \cite{lieberman2005}. The resulting current density was 200 $\mu$A cm$^{-2}$ at the sample. During operation, the sample temperature remained below 100 $^\circ$C. It is well established that the ion induced pattern formation on Si does not depend significantly on temperature between $\sim$20 $^\circ$C and $\sim$150 $^\circ$C \cite{SanchezNanotech}.
Commercially available epi-polished Si(100) wafers (n-type, $\sim$10 $\Omega\cdot$cm) were used as samples. The metal incorporation was switched on and off by using two different masks that also served to fix the samples on the sample holder. Metal atoms sputtered from a stainless steel mask may be re-deposited on the sample surface due to partial ionization in the plasma. Alternatively, the steel mask was covered by an evaporated layer of Si of $\sim$1 $\mu$m thickness, so that the metal incorporation is largely suppressed. For post-sputtering experiments, the Si(100) surfaces were pre-patterned by ion bombardment with the same parameters as stated above, removed from the chamber and re-mounted with the Si covered mask for subsequent irradiation at different fluences.
The surface topography was analyzed {\it ex situ} by atomic force microscopy (AFM) in tapping mode. The surface metal content was analyzed by Rutherford backscattering spectrometry (RBS) using a beam of 1.7 MeV He$^+$ ions at normal incidence. Ions scattered at an angle of $170^\circ$ were analyzed using a Si surface barrier detector.
\section{Results and discussion}
\subsection{Influence of metal incorporation on dot pattern formation}
\label{subsec:comparative study}
Different samples, which will be denoted as A and B in the following, were fixed with a bare stainless steel mask and the Si covered mask, respectively, as described above. They were irradiated during 15 minutes each, corresponding to an ion fluence of $1.1\times10^{18}$ cm$^{-2}$.
\begin{table}
\caption{\label{table1}Surface contents of Ar, Fe, and Cr as obtained from RBS measurements.}
\begin{indented}
\lineup
\item[]\begin{tabular}{@{}*{5}{cccc}}
\br
&&\centre{3}{Contents (10$^{15}$ cm$^{-2}$)}\\
\ns
&&\crule{3}\\
Sample&Mount&Ar&Fe&Cr\\
\mr
A&Stainless Steel&1.63$\pm$3\%&1.53$\pm$3\%&0.40$\pm$13\%\\
B&Si coverage&1.64$\pm$3\%&0.10$\pm$50\%&0.01$\pm$100\%\\
\br
\end{tabular}
\end{indented}
\end{table}
Table \ref{table1} shows the contents of three representative elements on both samples after irradiation. The Ar content is not influenced by the amount of the metal impurities while the amount of Fe and Cr is reduced on sample B by more than one order of magnitude. The relative amounts of Fe and Cr correspond to the atomic fraction of Cr in SS 304 ($\sim$18 \%). Figure \ref{fig:comparison} shows the power spectral density (PSD) functions obtained from the AFM images (inset) of the two samples. An ordered nanodot array is formed only in the condition of high metal incorporation, with a mean roughness of 0.85 nm. The presence of the peak in the PSD function reveals a characteristic periodicity of 50 nm. In contrast, no correlated structures are observed on sample B at a roughness of 0.23 nm.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.65\linewidth]{figure1}
\caption{AFM images (inset) and the corresponding PSD functions from Si(100) samples after sputtering in the presence and absence of metal incorporation (A and B, respectively). Two lines with slopes -2 and -4 are plotted for comparison (see text).}
\label{fig:comparison}
\end{center}
\end{figure}
Basically, the surface morphology evolution results from an interplay between surface instabilities and relaxation processes. According to Bradley and Harper (BH) \cite{BradleyHarper}, the morphology evolution can be described by a partial differential equation of the surface height $h(\mathbf{x},t)$ where $\mathbf{x}$ denotes the two-dimensional lateral surface coordinate and $t$ the time. At normal incidence, the laterally isotropic BH equation reads,
\begin{equation}
\label{eq:lin_eq}
\frac{\partial h(\mathbf{x},t)}{\partial t}=-v_0+\nu \nabla^2 h-D \nabla^4 h.
\end{equation}
$v_0$ is the erosion rate of the unperturbed planar surface. $\nu$ and $D$ are parameters which depend on the experimental conditions. The second term on the right hand side represents the curvature dependent roughening and smoothing for negative and positive values of $\nu$, respectively. Previous studies of low energy ion sputtering on Si surfaces suggest that the surface processes related to the second order derivative of the surface height are the curvature dependent erosion (CDE) rate \cite{BradleyHarper, SigmundJMS1973} and ballistic smoothing \cite{CarterPRBball, MoselerScience, FrostJPCMrev, VauthMayr2007}. The third term stands for a combination of relaxation processes while only ion-enhanced viscous flow (IVF) \cite{UmbachPRL} are relevant for low energy ion sputtering on Si surfaces at room temperature \cite{VauthMayr2007, ZiberiPRB}.
Equation \ref{eq:lin_eq} is linear and can be solved by Fourier Transformation, yielding for the (angular averaged) PSD function \cite{FrostJPCMrev, MakeevNIMB}
\begin{equation}
\label{eq:PSD}
\eqalign{\mathrm{P}(k,t)=\mathrm{P}(k,0)\exp(2r_kt),\\
r_k=-\nu k^2-Dk^4,}
\end{equation}
where $k$ denotes the spatial frequency, and $r_k$ the corresponding growth rate. In the case of $\nu<0$, $r_k$ has a positive maximum at $k_c=\sqrt{-\nu/2D}$, leading to a correlated pattern with the characteristic frequency $k_c$.
However, in contrast to the present results, the surface processes described above cannot directly be related to the presence of impurities at the surface. Shenoy \etal \cite{ShenoyPRL} have shown that differences in sputtering yields and in surface diffusivities of different surface components can lead to modulations in topography and composition. In connection with the present experiments, a compositional inhomogeneity is indicated by the cross-sectional high-resolution transmission electron microscopy image of the dot patterned Si(100) surface shown in figure \ref{fig:HRTEM}. The nanodots are crystalline bumps with an amorphous top layer of $\sim4$ nm. The height and the periodicity of the dots are $\sim5$ nm and $\sim50$ nm, respectively. The intensity contrast is attributed to an accumulation of the metal impurities on top of the bumps. For a quantitative estimation, the darkened volume of $\sim$2.5 nm height contains $\sim$1.25$\times$10$^{16}$ Si atoms cm$^{-2}$. The areal density of metals concentrated on the crests can be estimated to be 2-3 times larger than the average total metal areal density of about 2$\times$10$^{15}$ cm$^{-2}$ (see table 1). Thus, the composition is not inconsistent with that of a stoichiometric metal disilicide. This would confirm the conjecture by Ozaydin \etal \cite{OzaydinAPLMo} that Mo silicides might be present during their ion sputtering experiments on Si surfaces. For the present system, a local decrease of the sputtering yield at the contaminated areas may be anticipated due to the larger surface binding energy of the silicide compared to pure silicon. Binary collision computer simulations using the TRIDYN program \cite{Moeller1988, Moeller2001} confirm an decrease of the Si sputtering yield from FeSi$_2$ vs. Si by $\sim$ 10 \%. Thus, once the non-uniform contamination is established, the differences of the sputter yield will indeed promote the pattern formation. There might be additional collisional effects which influence the pattern formation under metal contamination. The transition from holes to dots found by Sanchez-Garcia \etal \cite{SanchezNanotech} was qualitatively attributed to a change in the shape of collision cascade due to the metal impurities in the near-surface region.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.6\linewidth]{figure2}
\caption{Cross-sectional HRTEM image of the correlated nanodots generated with metal incorporation. An amorphous top layer is seen which is darkened on top of the crests due to atomic number contrast.}
\label{fig:HRTEM}
\end{center}
\end{figure}
As an alternative or additional possible mechanism, local stress might influence the dot pattern formation during ion sputtering. Previous studies show that the stress generated in the near-surface region is large enough to influence the morphology evolution \cite{ChanChasonJVSTA, ChanChasonNIMB, Kalyanasundaram2006, Dahmen2003}. Recently, Ozaydin \etal \cite{OzaydinJVSTB} found a tensile stress which they attributed to Mo incorporation during ion sputtering. Such tensile stress could cause the growth of surface protrusions as a consequence of relaxation \cite{OzaydinJVSTB}. This might even be promoted by the non-uniform contamination described above. Nevertheless, it remains speculative if a stress mechanism could also contribute to the present findings.
Thus, there is no conclusive information on the role of the metal incorporation neither from previous studies nor the present one. On the basis of the results presented above, we conclude that metal incorporation induces a new surface instability in addition to the curvature dependent erosion rate. Assuming that this new instability can be expressed as well by a second order derivative term as $\nu_{\mathrm{m}} \nabla ^2 h$, we rewrite the growth coefficient $r_k=-(\nu_{\mathrm{CDE}}+\nu_{\mathrm{m}}+\nu_{\mathrm{b}}) k^2-Dk^4$, where $\nu_{\mathrm{CDE}}$ and $\nu_{\mathrm{b}}$ represent the curvature dependent erosion rate and ballistic smoothing, respectively. Correlated patterns are generated when $\nu_{\mathrm{CDE}}+\nu_{\mathrm{m}}+\nu_{\mathrm{b}}<0$ which means that the combination of the two instabilities dominate the $k^2$-dependent surface process and compete with $k^4$-dependent smoothing mechanisms. In the absence of metal incorporation no correlated pattern is formed on the surface, indicating that the instability induced by curvature dependent erosion rate is overwhelmed by the relaxation processes. It is worth to note that the PSD function of sample B shown in figure \ref{fig:comparison} exhibits a transition from a $k^{-2}$ to a $k^{-4}$ behavior at the spatial frequency of approximately 0.05 nm$^{-1}$, as indicated by the two gray lines. This observation is in agreement with the asymptotic PSD function in the case of surface smoothing ($r_k<0$ for the entire spectrum) \cite{FrostJPCMrev, MakeevNIMB}
\begin{equation}
\label{eq:asymp}
\mathrm{P}(k,t \rightarrow \infty)\propto r_k^{-1},
\end{equation}
where the PSD function is approximately proportional to $k^{-2}$ at small frequencies and $k^{-4}$ at large frequencies.
\subsection{Smoothing of the pre-patterned surface}
For further investigation of the smoothing mechanisms, patterned samples were produced by ion sputtering with metal contamination as described above, and subsequently irradiated without metal contamination.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.65\linewidth]{figure3a_3d}
\caption{1$\mu$m$\times$1$\mu$m AFM images of Si(100) surfaces pre-patterned by 1200-eV Ar$^+$ ions at normal incidence with a stainless steel mask (a) and post sputtered with a Si covered mask at fluences of $1.0\times10^{17}$ cm$^{-2}$ (b), $2.0\times10^{17}$ cm$^{-2}$ (c) and $3.0\times10^{17}$ cm$^{-2}$ (d).}
\label{fig:dotpattern3}
\end{center}
\vspace{0cm}
\end{figure}
Figure \ref{fig:dotpattern3} show the AFM images from the original surface (a), which is similar to inset A in Fig. 1, and after post-sputtering with three different fluences of $1.0\times10^{17}$ cm$^{-2}$, $2.0\times10^{17}$ cm$^{-2}$ and $3.0\times10^{17}$ cm$^{-2}$ ((b)-(d)), respectively. The evolution of the root mean square (rms) surface roughness obtained from 1 $\times$ 1 $\mu$m$^2$ AFM images and of the Fe surface content obtained from RBS are shown in figure \ref{fig:metalpreroughen} (a). The morphology evolution can be divided into two regimes: At fluences smaller than $1.0\times10^{17}$ cm$^{-2}$, the surface morphology does not visibly change, as shown by figure \ref{fig:dotpattern3} (a) and (b). The roughness remains almost constant and even slightly increases. At fluences larger than $1.0\times10^{17}$ cm$^{-2}$, the nanodot pattern gradually decays with the inter-dot distance being unchanged until the surface becomes flat. An exponential decay of the roughness is observed. The smoothing effect in this regime is also confirmed by figure \ref{fig:dotpattern3} (b)-(d).
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.65\linewidth]{figure4a_4b}
\caption{Post-sputtering of Si(100) surfaces pre-patterned by ion sputtering under metal incorporation. (a) Root mean square (rms) surface roughness $W$ (squares) and surface content of Fe (circles) versus fluence $\Phi$. The solid and dashed lines are plotted to guide the eyes. (b) PSD functions (calculated from 1 $\mu$m$\times$1 $\mu$m AFM measurement) of the samples post-sputtered with the indicated fluences.}
\label{fig:metalpreroughen}
\end{center}
\vspace{0cm}
\end{figure}
The non-monotonic behavior in morphology evolution is attributed to the surface metal content. Based on the observation that the surface metal content decreases with ion fluence (see Fig. 4 (a) for the Fe surface content), we assume that at the beginning of post-sputtering the metal atoms inherited from the pre-patterning process still remain on the sample surface although further metal incorporation is now prevented. Therefore, the surface instability is still large enough to compete with the smoothing mechanisms. Since the surface morphology of the pre-patterned samples has reached the saturation regime, the pattern does not change in the early stage of post-sputtering at fluences which are small compared to the pre-irradiation fluence. With post-sputtering being continued, the incorporated metal atoms are gradually removed from the surface, so that finally the smoothing mechanisms dominate the morphology evolution. In this sense, a threshold value of the metal content is expected for the roughening/smoothing transition. When the metal content is larger than this threshold value, the surface instability is large enough to trigger dot pattern formation. The present experiments indicate that this threshold value for a metal impurity of mass $\sim$ 55 amu in Si is between $0.6\times10^{15}$ cm$^{-2}$ and $1\times10^{15}$ cm$^{-2}$, for Ar ion irradiation under the present conditions.
Figure \ref{fig:metalpreroughen} (b) shows the evolution of the PSD function in the decay regime. At the first sight, we find smoothing over the entire spectrum with a rapid decrease of high frequencies and a successive decrease of lower frequencies. Finally, the morphology reaches a steady state. As expected, the steady-state PSD function shows a similar behavior as the one obtained after sputtering of a virgin sample without metal contamination (see Fig. 1, curve B). According to the discussion in section \ref{subsec:comparative study}, the transient from a $k^{-2}$ to a $k^{-4}$ behavior in the PSD function corresponds to the asymptotic solution (equation (\ref{eq:asymp})) in the case of surface smoothing. This is a strong indication that the relaxation processes include both $k^2$- and $k^4$-dependent smoothing mechanisms.
Surface smoothing during ion sputtering has been investigated in several studies. Vauth and Mayr \cite{VauthMayr2007} suggest that ion-enhanced viscous flow or ballistic smoothing dominates for high or low spatial frequencies, respectively, during keV ion smoothing of amorphous surfaces. Although Frost \etal \cite{FrostJPCMrev} observed only a $k^2$-dependent surface smoothing during ion sputtering on Si, they also found first evidence for a transition from a ballistic smoothing ($\propto k^2$) to viscous flow smoothing ($\propto k^4$) in the PSD function composed from two AFM images with different scanning size. IVF has been suggested in several studies as the dominant $k^4$-dependent smoothing mechanism during low-energy ion sputtering on Si surfaces \cite{VauthMayr2007, ZiberiPRB, AlkemadeAPL}. The concept of ballistic smoothing was firstly put forward by Carter and Vishnyakov \cite{CarterPRBball} as a result of the net displacement of the forward recoils moving preferentially along the ion beam direction. The ion-impact-induced downhill current mechanism proposed by Moseler \etal \cite{MoselerScience} on the base of molecular dynamics computer simulation is considered phenomenologically virtually indistinguishable from Carter and Vishnyakov's proposal \cite{DavidovitchPRB}. Zhou \etal \cite{HZhou2008} observed $k^2$-dependent smoothing on Al$_2$O$_3$ surfaces during Ar$^+$ ion erosion at normal incidence. They attribute this effect to the ballistic atomic downhill current \cite{MoselerScience}. Accordingly and based on the discussion above, we suggest that ballistic smoothing and IVF are the two dominant relaxation processes during normal incidence Ar$^+$ ion bombardment of Si surfaces.
\section{Conclusions}
In conclusion, with 1200-eV Ar$^+$ ions bombarding Si(100) surfaces at normal incidence and below 100 $^\circ$C, we have observed dot pattern formation and surface smoothing in the presence and absence of metal incorporation, respectively. Metal incorporation corresponding to less than one monolayer critically influences the roughening/smoothing behavior. In addition to the well-known curvature-dependent erosion rate, an additional surface instability is generated, which is consistent with laterally non-uniform sputtering due to preferential accumulation of metal atoms at the nanodots. In the case of smoothing, a transition from a $k^{-2}$ to a $k^{-4}$ behavior is observed in the asymptotic PSD function, which is ascribed to ballistic smoothing and ion-enhanced viscous flow as the dominant surface relaxation mechanisms.
\ack
The authors acknowledge the experimental assistance of Dr. R. Gr\"{o}tzschel, Dr. A. Rogozin, Dr. B. Schmidt, Dr. A. M\"{u}cklich and D. Hanf. This work has been supported by DFG FOR 845 and by NSFC through Grant No. 60638010 and No. 60776038. The life cost of Jing Zhou's research stay in Dresden is supported by China Scholarship Council.
\section*{References}
|
\section{Introduction}
Richard Dawkins, who first introduced the term \emph{meme}, described it as a replicator: a unit of cultural transmission which propagate itself in the meme pool in analogy to what genes do in the gene pool \citep[Chap. 11]{Dawkins1976}. More generally, he proposed a deep analogy between biological and cultural evolution, with the latter encompassing memes that replicate trough imitation, mutate when new cultural variants are produced and compete for human brain resources. While this analogy must not be taken too far --- critics underline, for instance, that memes have often weakly-defined boundaries, replicate with much less fidelity than genes or that selection pressures on memes derive from little understood mechanisms \citep[e.g.][]{Atran2001,RB2005,Wimsatt1999} --- at a more general level there are little doubts that some kind of evolutionary process underlies culture development and change \citep{Boyd1985,Dawkins1976,Mesoudi2004,Plotkin1993,Rogers2008,RB2005}.
Thinking to culture as an evolutionary process is not without problems. Among them, especially intriguing is the fact that, given a constant or sufficiently slow changing selective pressure, replicators should converge in the long run into a small number of highly adapted forms. However, as an enormous number of genera and species live side to side in the biological world, cultural variants coexists in the cultural world. The paradox of the presence of a large biodiversity, resulting from apparently converging selective pressures, is resolved in natural sciences using the concept of niche adaptation and taking into account the constant moving of species between regions \citep{Hubbell2001,Sugihara2003,Tilman2004}. However, few social scientists have taken the problem from this point of view, focusing instead their research on mechanisms connected to some structural characteristics of human societies rather than with the cultural environment itself \citep[e.g.][]{Axelrod1997a}.
We present here a model of cultural evolution where agents are able to exploit different niches on an artificial environment. This ability is strictly linked with their belonging to a memetic group (that can be viewed as a culture). More precisely, agents belonging to a group share a part of their ``memotype'' \citep{Speel1997}. Instead of imitating any other agent in the population, agents updating their strategies tend follow the ones already present in their own group. However, agents can also choose to change their group: a fact that produces a multilevel selection process. Note that, while multilevel selection is a source of unending debate in biology \citep[for a recent review see][]{Kohn2008}, it may be more plausible in reference to cultural evolution \citep{Retal2003,RB2005,Soltis1995,WS1994}. In our model, belonging to a specific memetic group allows agents to exploit different niches of their environment with, as a result, the coexistence of different memotypes in the same population. This in contrast with a model where selection takes place only at the level of individual agent leading instead to the fixation of a single memotype. Finally, we show that a situation when changes in the meme pool of agents also modify the environment where they interact favor the establishment of coexisting memetic groups in the system and, more generally, it is able to model realistic evolutionary processes.
The remaining of the paper is organized as follows. Section \ref{theory} illustrates the research background. Section \ref{model} defines our cultural evolution model and presents the corresponding results. Section \ref{coevolution} defines and presents the results of the niche construction model, where agents' evolution affects also the state of the environment. Finally, Section \ref{discussion} discusses the results.
\section{Groups, niches and cultural differences}\label{theory}
One of the seminal and most well known researches trying to model culture evolution and dissemination is Robert \citet{Axelrod1997a} work on \emph{Social influence}. Its main result is that cultural differences can be sustained, despite the effect of mechanisms like homophily and social influence, by the creation of boundaries between ``cultural regions''. In Axelrod's model, $n$ agents interact on a regular lattice, typically a 2D, $10 \times 10$ grid. Each agent has a cultural state defined by a vector of $F$ elements (cultural features), which can assume any of $q$ different discrete values. Agents' initial state is defined randomly. Subsequently, a pair of neighboring agents is selected in each time step, with their interaction probability depending on the respective states: it is zero when the couple shares no features and increases proportionally to the number of shared features (homophily assumption). When two agents interacts, one of them copies one of the unshared features of the other (social influence). The simulation continues until a stable state is reached, either because all agents reach the same state or because stable boundaries emerge between cultural regions composed by agents sharing no features. Axelrod shows that, under a wide range of condition, the latter outcome is more likely: a fact that tend to maintain over time cultural differences in the system.
This result is interesting, especially because it holds despite the fact that the interactions occur in a regular space presenting no possibilities of niche differentiation and that the modeled features possess no adaptive meaning for the agents. The Social influence model represents hence an alternative to biological explanations of biodiversity based on the niche adaptation concept. Unfortunately, subsequent works have shown that Axelrod's results depend critically on a few implausible assumptions, including the absence of noise in the transmission process \citep{Klemm2003a} and discrete cultural variants \citep{Flache2006}, possibly drawn from a large set \citep{Klemm2003}. Moreover, Axelrod's results hold only under the assumption of \emph{perfect homophily}, i.e. a null interaction probability between agents sharing no cultural features. The creation of impermeable boundaries represents indeed the key factor able to guarantee pluralism \citep{Axelrod1997a}. The problem with this assumption is that no boundary is perfectly impenetrable in the real world, while relaxing it leads inevitably to an homogeneous state in the long run.\footnote{\citet{Flache2006} first discussed this fragility of the model, even if they apparently failed to recognize its full power in undermining Axelrod's conclusions.}
The example in Figure \ref{fig:epsilon} illustrates the effect of introducing an arbitrarily small probability of interaction between agents sharing no features ($\varepsilon$ ) in a system where six agents interact on a one-dimensional lattice.\footnote{A 1D system is used here for simplicity, but the same argument holds for multidimensional systems as well, including the regular 2D lattice used by Axelrod.} Each agent has five cultural features that can assume any integer value between zero and nine. Line (a) presents an equilibrium situation as described by Axelrod, with a stable cultural boundary between the third and the fourth agent. Line (b) shows the same situation but for a small $\varepsilon$ probability of interaction between the two middle agents sharing no cultural trait. The probability that the fourth agent is selected and copies one feature from the third one is $\varepsilon / 2n$. Taking into account that also the third agent has the same probability of copying one trait from the fourth one, the overall interaction probability is $\varepsilon / n$. No matter how small, this probability is strictly greater than zero and, given a sufficient amount of time, the fourth agent will actually copy one feature from the third one (or vice versa). This leads to the situation depicted in line (c), with an increase of the interaction probability between the fourth agent and \emph{both} its neighbors. More generally, the system moves away from the stable state in a process that is self-reinforcing. Actually it is quite intuitive that a system with a probability of interaction between agents sharing no features is greater than zero possesses only one stable state: the one in which all agents are identical. Simulations based on the Axelrod's model confirm this intuition.
\begin{figure}[t]%
\centering
\fbox{\parbox{\linewidth}{\centering
\begin{tabular}{lccccccccccc}
(a) & 01234 & (0) & 01234 & (0) & 01234 & (0) & 56789 & (0) & 56789 & (0) & 56789 \\
&&&&&&&&&&&\\
(b) & 01234 & (0) & 01234 & (0) & 01234 & ($\varepsilon / n$) & 56789 & (0) & 56789 & (0) & 56789 \\
&&&&&&&&&&&\\
(c) & 01234 & (0) & 01234 & (0) & 01234 & ($0.2/n$) & 06789 & ($0.8/n$) & 56789 & (0) & 56789 \\
\end{tabular}
}}
\caption{Effect of the introduction of an arbitrarily small probability of interaction between agents with no traits in common in a 1D version of Axelrod's Social influence model. The probability of interaction between two neighbors is in parenthesis.}
\label{fig:epsilon}
\end{figure}
Our argument is that Axelrod's model most fundamental problem derives from the uniform space where agent interacts, which does not allow the exploitation of different niches by different cultures. This is vastly different from what happens for biological evolution, and also from what we know from the functioning of social systems, where different groups are usually able to find vastly different ways of making a living by exploiting different resources. For instance, pastoral communities can live close to farmers by using different natural resources. Moreover, higher level specialists, like merchants, can rely only indirectly on natural resources by exploiting niches built up by the work of other individuals.
The link between different environments (or niches) and different cultures is supported by some classical anthropological studies \citep{Harris1979}. \citet{Henrich1998} developed a model where different subpopulations are able to evolve specific adaptations to the different environments where they live, despite a significant amount of migration among them. The core of the model is ``conformist transmission'', i.e. the fact that agents tend to imitate disproportionately the most commons behavior in their group. From our point of view, it is especially important to remark that conformist transmission, while it favors within-group cultural homogeneity, appears to be able to maintain between-group differences. However, the model achieves this result by building separate environments with opposite adaptive needs. It does hence not explain how cultural diversity can be maintained when agents interact \emph{in the same environment}. On the contrary, the effect of conformist transmission in a single environment setting is likely to lead to a homogeneous state.
An alternative approach is thinking about cultures in terms of biological species \citep{Mace2005, PM2004}. The core of the species concept is that organisms from different species cannot interbreed successfully. Biodiversity derives hence from niche adaptation, but it is maintained thanks to strong barriers to the gene flow. If the analogy between cultures and species holds, it implies that the ``interbreeding'' among cultures must be limited in order to maintain diversity. This is actually what happens in human groups, where a number of mechanisms exist to limit cultural and possibly genetic influences from other groups \citep{PM2004,RB2001}. While this may surprise cosmopolitan citizens of modern cities, empirical studies actually show that a large proportion of cultural transmission occurs phylogenetically within groups, including (of course) families and (more generally) homogeneous cultural groups \citep{Guglielmino1995,Mace2005}.
The consequence of a reduction the meme flow among cultures is a decrease of within-group variation and a corresponding increase of between-group variation: a situation that increases the possibility that selection processes occur also at the group level, as acknowledged also by group-selection skeptics \citep[e.g.][Chap. 13]{Dawkins1989}. This is actually one of the reasons why group selection processes are more likely to be a significant force in cultural than in biological evolution \citep{Retal2003,RB2005,Soltis1995,WS1994}. Another reason is that, although appealing, the cultures-as-species metaphor is wrong in one important detail: unlike biological organisms, that are rarely able to acquire DNA from other living beings,\footnote{This happens mainly in bacteria and other microbes: asexual organisms for whom species are weakly defined.} memes in human brains are relatively easy to change at any point of an individual's life. This has important consequences, because intergroup migration represents a major factor undermining group-selection processes in biological evolution \citep{MaynardSmith1976}. On the contrary, human migrants often acquire the culture of their new groups, making hence migration less effective in reducing between-group variation. This, along with the high levels of intergroup conflict typical of human societies \citep[e.g.][]{Soltis1995}, makes group selection processes plausible, if not probable, for cultural evolution \citep[Chap. 6]{RB2005}: a fact that justifies our choice of modeling a multilevel selection process (see below).
Before proceeding with the model definition, one more point should be carefully discussed: the value of a given culture for the people who share it. Using a strict ``meme-eye view'' the only interest of the memes would be to create as many copies of themselves as possible. However, from the point of view of individuals hosting these memes, their value is clearly linked to the opportunities that they offer in finding adequate ways of making a living. Similarly, the adaptiveness of a given group to its environment depends on the culture shared by its members. Cultures allowing groups to exploit some environmental niche hence possess the value of a shared resource. While natural shared resources often present ``tragedy of the commons'' problems leading to their depletion or destruction \citep{Hardin1968}, cultural commons tend instead to increase their value with use \citep{Boiller2007,Hess2007}. For instance, the value of a software or of a piece of scientific knowledge increases as more people use it. The problem here is indeed not overuse, like in natural commons, but reaching the critical mass that makes a given culture self-sustaining.
In order to coexist, cultures should also allow their members to exploit different niches of the social or of the natural environment (or both). Otherwise, in absence of perfectly impermeable boundaries, selection processes tend to lead to the fixation of a single culture. Each group must hence be able to define and use its own ``cultural commons'', i.e. the shared cultural resource allowing the exploitation of a specific niche in its environment. Note that, even if cultural commons do not suffer from overuse, still they have to be maintained and protected from ``erosion'', i.e. the undermining of their internal coherence following uncontrolled changes. However, no sustainable use of a commons is likely without the definition of boundaries, including limits of its users' group \citep{Ostrom1990}. This point parallels hence the above group-selection argument in supporting a strong influence of group barriers in making the coexistence of different cultures possible.
Our arguments can be summarized in two hypotheses. First any coexistence among cultures need boundaries. They may not be perfectly impermeable, as in Axelrod's model, but still they should be strong enough to permit the definition of in-group and out-group members. Most of the times, boundaries in human societies are quite strong and produce within-group cooperation and favoritism and between-group competition or conflict \citep{Bernhard-etal2006,Iida2007,RB2001,Soltis1995,Tajfel1986}. Nevertheless, boundaries are not a sufficient condition since, in absence of the possibility of exploiting different niches, competition will lead to the disappearance of the weakest groups and, in the long run, to the fixation of a single culture. Our second hypothesis is hence that cultures can coexist by allowing their members to exploit different niches of their social-ecological system,\footnote{For a definition of social-ecological systems, see \citet{Berkes1998}.} just as specialization (and often speciation) allows different organisms to exploit different ecosystem niches. The models presented in the next sections will explore these ideas.
\section{The cultural evolution models}\label{model}
In order to test our hypotheses, we designed two simulation models. The first one, named \emph{Base}, represents our benchmark. It defines our artificial agents and environment, but supports no group processes or boundaries. In the \emph{Group} model, boundaries among groups of agents are instead explicitly defined, even if they are not perfectly impermeable.
\subsection{\emph{Base} model definition}
The base model is formed by $N$ agents, representing organisms, interacting on an abstract environment.\footnote{All models have been implemented in C++. Codes are available upon request.} The environment $E$ is simply a sequence of $n$ binary numbers. It is randomly determined at the beginning of each run and remains subsequently constant. Each of the $N$ agents possess a ``memotype'' $M$ formed by a sequence of $m = 6$ binary numbers. This gives a total of 64 possible memotypes. Adopting a population of 256 agents, we have an average of 4 agents per memotype. At the beginning of each run, memotypes are determined randomly, while they are subsequently free to evolve. During each period of the game agents earn a variable number of points. The payoff of each agent depends on the correspondence between its memotype and the environment. More specifically, each time that $M$ corresponds exactly to a sub-sequence of $E$ the agent earns one point (Fig. \ref{fig:example} left). The period payoff is simply the sum of all points earned by a given agent in a single period.
\begin{figure}[t]%
\fbox{
\parbox{\linewidth}{\large
\[
\begin{array}{lc@{}c@{}c@{\hspace{1cm}}c@{}c@{}c}
& & \text{\normalsize \emph{Base} model} & & & \text{\normalsize \emph{Group} model} & \\
\noalign{\vskip3pt}
\text{\normalsize Memotype} & & \overbrace{101000}^{M} & & & \overbrace{101}^{M_g}\overbrace{000}^{M_i} &\\
\text{\normalsize Environment} & \ldots 01011 & \overbrace{101000}^{} & 11011 \ldots & \ldots 01011 & \overbrace{101000}^{} & 11011 \ldots
\end{array}
\]
}}
\caption{Example of memotype-environment correspondence in the \emph{Base} and the \emph{Group} models.}%
\label{fig:example}%
\end{figure}
A fundamental feature of the model is the imitation process occurring at the end of each period. During imitation, one of the agents with the lowest period payoff copies the memotype of one of the agents with the highest period payoff. However, a small mutation probability $\mu$ exists. When a mutation occurs the new memotype of the low payoff earner is randomly defined. Note that the term mutation should not be taken here in a strict biological meaning. Rather, it stands for innovations and new ideas entering the game and allowing for the exploration of new parts of the environment.
Before proceeding with the definition of the \emph{Group} model, it is important to understand how the length of the environment influences the number of niches where agents can make their living. We define a niche as a sub-sequence of $E$ of length $m$ leading to the maximal period payoff for an agent holding the corresponding memotype. For instance, let us suppose the case where $E = \{0,0,0,1,1,1\}$ and $m=2$. Both the memotypes $M_{00}=\{0,0\}$ and $M_{11}=\{1,1\}$ earn two points per period, while $M_{01}=\{0,1\}$ earns one point and $M_{10}=\{1,0\}$ zero points. The two memotypes $M_{00}$ and $M_{11}$represent cultural variations that are neutral in respect to the selection process, both leading to the maximal payoff. Following the definition above, $E$ possesses hence two niches.
In order to choose the environment length to use in the model, we explored the number of niches in our environment for $m=6$ and $n \in \{10,15,\ldots,300\}$. The results of a sample of 10000 environments for each value of $n$ are reported in Figure \ref{fig:environment}. It is clear that the number of niches (but also the standard deviation) reach a peak at $n \approx 3m$, reduces rapidly up to $n \approx 100$, while for higher values of $n$ it shows a continuous, although extremely slow, decline .\footnote{Note that a similar trend of peak as subsequently slow decline is present also for other values of $m$.}
\begin{figure}[t]%
\centering
\includegraphics[width=0.45\textwidth]{environment}%
\caption{Average number of niches per environment lengt for a memotype of length $m=6$. Standard deviations are represented as dotted lines.}%
\label{fig:environment}%
\end{figure}
It would be tempting to choose for our simulations values of $n$ leading to a high number of niches. However, the standard deviation corresponding to low values of $n$ is extremely high (Fig. \ref{fig:environment}), a factor that could undermine the stability of our results. We hence preferred to use the smaller value of $n$ leading reasonably close to the level of 2 niches. This resulted to be $n=150$, corresponding to an average of 1.97 niches per environment.
\subsection{\emph{Group} model definition}
The \emph{Group} model introduces boundaries in the game, making some memes easier to adopt than others. More specifically, the \emph{Group} model maintains all the elements of the \emph{Base} one except for the fact that the memotype of each agent is now formed by the concatenation of two sequences, $M_g$ and $M_i$, composed by $m_g$ and $(m - m_g)$ binary numbers respectively. The former identifies the section of the memotype common to each group while the latter defines the agent's specific memotype. Once the two sequences are concatenated, the memotype-environment correspondence is checked as in the \emph{Base} model (Fig. \ref{fig:example} right). The imitation process occurs as in the previous case, but it now concerns only $M_i$. Moreover, imitators and imitation targets are selected inside each group. This means that, for each $M_g$ sequence, one hight-earner agent and one low-earner agent are selected, with the latter copying the former's $M_i$. The mutation probability is as in the \emph{Base} model.
The \emph{Group} model introduces a further process, called ``migration'', working side to side with imitation. Migration regards only $M_g$ and concerns the whole population of agents. At the end of each period, one of the agents with the lowest payoff is selected as ``migrant''. The ``migrant'' changes its group and adopts the $M_g$ sequence of one of the agents with the highest payoff. Again a mutation probability $\mu$ exists, modeling the individual decision to form a new group. Note that making the mutation probability of migration different from the imitation one proved not to be crucial in influencing the final result in some pilot runs. In order to limit the number of parameter in the simulation, we hence decided to maintain a single parameter for both kind of mutations.
\subsection{Results}\label{results}
We explored three different values of the mutation probability parameter and three dimensions of $m_g$, namely $\mu \in \{0.01,0.05,0.10\}$ and $m_g \in \{2,3,4\}$. In order to take into account the stochastic elements of the models, we performed 1000 runs for each parameter configuration. We hence tested 3 parameter configurations for the \emph{Base} model and $3 \times 3$ parameter configurations for the \emph{Group} one, for a total of 12000 runs. Each run lasted for 5000 periods and included 256 agents.
Figure \ref{fig:typical} presents the plots resulting from typical runs of the \emph{Base} and \emph{Group} models under different parameter conditions. It is clear than, while the \emph{Base} model tends to produce a single large memetic group (i.e. a group of agents sharing the same memotype), the \emph{Group} one is more favorable to the co-existence of different ``cultures''. However, this holds mainly for low values of $\mu$, while higher ones tend to drive the \emph{Group} model close to the results of the \emph{Base} one.
\begin{figure}[p]%
\centering
\includegraphics[width=.325\textwidth]{base01} \includegraphics[width=.325\textwidth]{base05}
\includegraphics[width=.325\textwidth]{base10}\\
\includegraphics[width=.325\textwidth]{group2-01} \includegraphics[width=.325\textwidth]{group2-05}
\includegraphics[width=.325\textwidth]{group2-10}\\
\includegraphics[width=.325\textwidth]{group3-01} \includegraphics[width=.325\textwidth]{group3-05}
\includegraphics[width=.325\textwidth]{group3-10}\\
\includegraphics[width=.325\textwidth]{group4-01} \includegraphics[width=.325\textwidth]{group4-05}
\includegraphics[width=.325\textwidth]{group4-10}\\
\caption{Agents in each memotype in typical runs of the \emph{Base} and the \emph{Group} models. The line width is proportional to the dimension of the memetic group.}%
\label{fig:typical}%
\end{figure}
In order to test whether the differences highlighted in the plots are statistically relevant, we counted the number of memetic groups in each of the 1000 runs performed for each parameter configuration. More specifically, we counted the number of memetic groups composed by more than 4 agent --- i.e a group larger than what we should expect if the agents' memotypes were randomly distributed --- and composed by more than 25 agents, --- i.e. a group including more than 10\% of the agents. A summary of the results is reported in Table \ref{tab:stats140}. Since at the beginning of each run the agents are randomly distributed among memotypes, the reported results refer only to the second half of the simulation, namely from period 2500 onwards.
\begin{table}[t]%
\centering
\begin{tabular}{cccr@{ }lr@{ }lr@{ }l}
\toprule
& & \multicolumn{7}{c}{Model}\\
\cmidrule(lr){3-9}
Memetic groups & $\mu$ & \emph{Base} & \multicolumn{2}{c}{\emph{Group}: $m_g=2$} & \multicolumn{2}{c}{\emph{Group}: $m_g=3$} & \multicolumn{2}{c}{\emph{Group}: $m_g=4$} \\
\midrule
$>4$ & 0.01 & 1.76 (1.20) & 3.20 (0.77) & $^{***}$ & 4.12 (1.65)& $^{***}$ & 3.66 (2.38) & $^{***}$ \\
$>4$ & 0.05 & 1.84 (1.26) & 2.03 (0.81) & $^{***}$ & 1.80 (0.88) & & 1.77 (0.97) & \\
$>4$ & 0.10 & 1.88 (1.39) & 1.52 (0.61) & $^{***}$ & 1.39 (0.55) & $^{***}$ & 1.35 (0.57) & $^{***}$ \\
\midrule
$>25$ & 0.01 & 1.56 (0.82) & 2.70 (0.75) & $^{***}$ & 2.12 (0.94) & $^{***}$ & 1.67 (0.87) & $^{**}$ \\
$>25$ & 0.05 & 1.61 (0.84) & 1.68 (0.67) & $^{*}$ & 1.44 (0.62) & $^{***}$ & 1.40 (0.58) & $^{***}$ \\
$>25$ & 0.10 & 1.62 (0.90) & 1.36 (0.51) & $^{***}$ & 1.28 (0.46) & $^{***}$ & 1.22 (0.41) & $^{***}$ \\
\bottomrule
\end{tabular}
\caption{Number of memetic groups including more than 4 and 25 agents in the final 2500 periods of the simulation. Results are averaged over 1000 runs per experimental condition with standard deviations in parenthesis. Significance codes ($t$ test between corresponding conditions of the \emph{Base} and the \emph{Group} models): $^{***}$ $p<0.001$, $^{**}$ $p<0.01$, $^{*}$ $p<0.05$.}
\label{tab:stats140}
\end{table}
The quantitative analysis confirms the hints offered by Figure \ref{fig:typical}. The \emph{Group} model leads to a significantly higher number of coexisting memetic groups than the \emph{Base} one, at least for $\mu=0.01$. A $t$ test shows indeed that, for all $m_g$ values, the number of memetic groups is significantly higher in the \emph{Group} model than in the corresponding \emph{Base} model. The standard for the \emph{Group} model is indeed the coexistence of two large groups (more than 25 agents) plus possibly one or two smaller groups. On the other hand, only one large group is almost ever present int the \emph{Base} model, leaving at most the space for one further smaller group.
The picture changes when $\mu$ increases. From one hand, the number of memetic groups in the \emph{Base} model tends to increase: a fact that is rather surprising, since high mutation rates weaken the ``efficiency'' of the selection process, leaving more room for alternative memotypes. On the other hand, increasing the mutation probability has rather dramatic effects for the \emph{Group} model. For any value of $\mu \geq 0.05$ the number of memetic groups in the \emph{Group} model is no longer higher than the one in the \emph{Base model}. For $\mu = 0.10$ this figure becomes even smaller than the corresponding one in the \emph{Base} model.
\section{The niche construction models}\label{coevolution}
The models presented above offer some interesting insights, but only partially answer to our questions. This for two main reasons. First, while they succeed in modeling the selection process, they do not reproduce the death and birth of groups typical of the cultural world. In other words, they do not lead to a real evolution of alternative cultures. Second, the \emph{Group} model lead to the coexistence of different memotypes only under the condition of high copy fidelity of cultural traits: something that is likely not to happen in the real world \citep{RB2005}.
What is probably missing from these models is that organisms often modify the environment where they live, affecting also the possibilities of making a living of other organisms. Some authors further expand this idea by arguing that, since the action of any organism modifies the availability of resources for other organisms, this changes the selective pressures existing in the environment. This process is called ``niche construction'' \citep{Laland2008,Odling-Smee2003} and there is no reason to imagine that it does not apply to cultural evolution. On the contrary, modern human economies represent a clear example of enormous differentiation based on what other individuals do, think, produce and consume. In order to reproduce this and similar facts, we hence decided to build two further models where the environment is no longer fixed, but varies throughout the simulation as a function of the agents' memotypes.
\subsection{Model definiton}
The new models derive from the previous ones and maintain their structure. More specifically, everything but the environment definition remains as in the \emph{Base} and the \emph{Group} models. We hence called the two new models \emph{NC-base} and \emph{NC-group}, where ``NC'' stands for niche construction. What changes is that the environment definition is no longer a process on its own. The environment is instead built by concatenating the complements of each element of the agents' memotypes. An example will make this point clearer. Given a model including three agents with memotypes $M_1 =\{0,0,0\}$, $M_2=\{1,1,1\}$ and $M_3=\{1,0,1\}$, the environment is $E=\{1,1,1,0,0,0,0,1,0\}$, which leads to a payoff of two, one and zero points respectively for the three agents. The choice of using the memotype complements permits to avoid the tautological condition where agents are adapted to themselves. On the contrary, each agent creates the possibility of making a living (a niche) for a different agent in a dynamic process that mimics what happens in the natural and the social world.
\subsection{Results}
In order to test the new models we used the same parameter configurations explored for the cultural evolution models (except, obviously, for the environment length which is now given by the number of agents times the memotype length). We hence performed a total of 12000 runs of the new models. Figure \ref{fig:niches} presents the plots of typical runs of the \emph{NC-base} and \emph{NC-group} models under different parameter conditions while Table \ref{tab:statsNiches} summarizes the statistics.
\begin{figure}[p]%
\centering
\includegraphics[width=.325\textwidth]{NCbase01} \includegraphics[width=.325\textwidth]{NCbase05}
\includegraphics[width=.325\textwidth]{NCbase10}\\
\includegraphics[width=.325\textwidth]{NCgroup2-01} \includegraphics[width=.325\textwidth]{NCgroup2-05}
\includegraphics[width=.325\textwidth]{NCgroup2-10}\\
\includegraphics[width=.325\textwidth]{NCgroup3-01} \includegraphics[width=.325\textwidth]{NCgroup3-05}
\includegraphics[width=.325\textwidth]{NCgroup3-10}\\
\includegraphics[width=.325\textwidth]{NCgroup4-01} \includegraphics[width=.325\textwidth]{NCgroup4-05}
\includegraphics[width=.325\textwidth]{NCgroup4-10}\\
\caption{Agents in each memotype in typical runs of the \emph{NC-base} and the \emph{NC-group} models. The line width is proportional the dimension of the memetic group.}%
\label{fig:niches}%
\end{figure}
\begin{table}[t]%
\centering
\begin{tabular}{cccr@{ }lr@{ }lr@{ }l}
\toprule
& & \multicolumn{7}{c}{Model}\\
\cmidrule(lr){3-9}
Memetic groups & $\mu$ & \emph{NC-base} & \multicolumn{2}{c}{\emph{NC-gr.}: $m_g=2$} & \multicolumn{2}{c}{\emph{NC-gr.}: $m_g=3$} & \multicolumn{2}{c}{\emph{NC-gr.}: $m_g=4$} \\
\midrule
$>4$ & 0.01 & 1.31 (0.66) & 2.46 (0.73) & $^{***}$ & 2.57 (1.03) & $^{***}$ & 2.70 (1.11) & $^{***}$ \\
$>4$ & 0.05 & 1.30 (0.62) & 2.02 (0.25) & $^{***}$ & 2.00 (0.31) & $^{***}$ & 2.16 (0.41) & $^{***}$ \\
$>4$ & 0.10 & 1.30 (0.65) & 2.12 (0.08) & $^{***}$ & 1.92 (0.25) & $^{***}$ & 2.02 (0.15) & $^{***}$ \\
\midrule
$>25$ & 0.01 & 1.25 (0.55) & 2.25 (0.57) & $^{***}$ & 2.13 (0.61) & $^{***}$ & 2.14 (0.45) & $^{***}$ \\
$>25$ & 0.05 & 1.23 (0.49) & 1.99 (0.19) & $^{***}$ & 1.92 (0.31) & $^{***}$ & 2.03 (0.21) & $^{***}$ \\
$>25$ & 0.10 & 1.23 (0.53) & 2.00 (0.02) & $^{***}$ & 1.88 (0.29) & $^{***}$ & 2.00 (0.13) & $^{***}$ \\
\bottomrule
\end{tabular}
\caption{Number of memetic groups including more than 4 and 25 agents in the final 2500 periods of the simulation. Results are averaged over 1000 runs per experimental condition with standard deviations in parenthesis. Significance codes ($t$ test between corresponding conditions of the \emph{NC-base} and the \emph{NC-group} models): $^{***}$ $p<0.001$, $^{**}$ $p<0.01$, $^{*}$ $p<0.05$.}
\label{tab:statsNiches}
\end{table}
From Figure \ref{fig:niches} two main consideration arise. First, the \emph{NC-group} model leads to the coexistence of memetic groups under all the examined parameter conditions and it is much less weakened by high mutation rates than the \emph{Group} model. Second, the evolutionary process looks more credible, with births and deaths of ``cultures'' and, more generally, with a greater number of changes during the course of the run.
Table \ref{tab:statsNiches} confirms that the number of coexisting memotype groups is significantly larger in the \emph{NC-group} model than in the \emph{NC-base} for all parameter configurations. Note also that standard deviations are much lower than in the cultural evolution models, witnessing a greater convergence of the niche construction models towards a clear equilibrium.
\section{Discussion}\label{discussion}
The main result of research is that group definition is effective in producing cultural heterogeneity, even in condition of a partial permeability of the boundaries. This support our first hypothesis, i.e. that the coexistence among cultures needs (imperfect) boundaries, under less restrictive condition than the Axelrod's model. In the \emph{Base} model coexistence cannot last, since selective pressures drive the whole population towards a single memotype even when alternative niches exists in the environment. The \emph{Group} model limits this process by reducing the flux between different memetic groups. This moves part of the selective pressure within the boundaries of each group, even if imitation and migration tend still to favor the best memotypes. The creation of group boundaries reduces within-group variation and increases between-group variation, leading hence to the maintaining of coexisting cultures. It is worth noting that this result supports the cultures-as-species idea advanced by \citet{PM2004}. In order to have cultural differences the memetic flux among groups must be limited, in analogy with the interbreeding limits existing among sexual reproducing organisms that stop the genetic flux and permits speciation.
A second point is that the group section of the memotype ($M_g$) has the value of a common resource for each group. Given the model definition, $M_g$ does not represent an arbitrary tag \citep[as, for instance, in][]{Axelrod1997a}, but possesses an actual adaptive value.\footnote{It is worth noting that some explorations of the model using purely symbolic group definition traits --- i.e. traits that were not checked agaisnt the environment --- did not produce a significant coexistence of cultures. While this is not crucial for the current discussion, it may represent the subject of further research.} Being a part of the memotype checked against the environment, $M_g$ contributes indeed actively to the agents' earnings. Successful group are hence groups with a group memotype well adapted to the current environment. This is equivalent to holding a specific knowledge or capacity in human societies. For instance, Renaissance artisans' guilds often kept secret their techniques in order to preserve the competitive advantage of their members. These techniques both helped to define the guilds themselves and represented the foundations of their wealth. Similarly, in our model $M_g$ both identifies the group and allows a successful exploitation of the environment. It hence represents an actual ``cultural commons'' for the group as a whole \citep[see][]{Hess2008}.
A final points regards the advantage of having coexisting cultures, which is mainly linked to the fact that each of them can exploit a different environmental niche.\footnote{Note that, in the real world, a dark side of cultural diversity exists, namely that interactions among different groups and cultures tend often to degenerate into conflicts \citep{RB2001,Soltis1995}.} The ``environment'' used in our model is too simple to create many niches. As a consequence, few memetic groups can be supported at the same time and instability, e.g. due to high mutation levels, quickly produces a convergence of the agent's memotypes (at least in the cultural evolution models). Nevertheless, in the \emph{Group} model two niches are, on average, present after every environment redefinition, which are promptly exploited by agents, at least under certain parameter conditions. What we found is actually a system where two different ``cultural commons'' (i.e. two different $M_g$) allow the persistence of two different memetic groups in a two-niche environment. This supports our second hypothesis, arguing that cultures coexistence is linked with the exploitation of different niches in a given social-ecological system.
The situation is somewhat more complicated in the niche construction models. A randomly defined environment of length $N \times m=1536$ holds, on average, only 1.35 niches: a value probably too small to support the coexistence of different memetic groups. This contrasts with our finding of at least two coexisting groups for all parameter configurations in the \emph{NC-group} model. However, the environment defined in the model depends on the agents memotypes. It is hence not random and can support more niches. Any memotype in the \emph{NC-group} model creates at least a niche for its complement. For instance, the memotype $M_1 = \{0,0,0\}$ creates a niche for the memotype $M_2 = \{1,1,1\}$ and vice-versa. The point here is that any large group creates a niche for a complementary group, building hence a two-niche environment. The observed convergence towards the coexistence of two large groups in the model should therefore not represent a surprise. It is actually the logical consequence of the niche construction process.
While more complex environments may lead to the coexistence of more groups, our model shows that the coexistence itself is possible given that (i) the environment possesses a sufficient number of niches and (ii) some process operates in order to keep the group boundaries at least partially impermeable (although it is not necessary to have perfect endogeneity as in Axelrod's model). This conditions can be checked against empirical evidence. For instance, \citet{Rogers2008}, in a research on the evolution over time and space of the shape of Polynesian canoes, show that functional features change slower than the symbolic ones. This is consistent with our arguments. The adaptiveness of functional features depends indeed on the law of fluid dynamics and on the technological abilities of Polynesian people. Since these two factors vary little across islands, it is likely that only one niche exists for functional features. The fact that we observe a convergence towards a single form --- carefully conserved over time and changing only as the result of occasional innovations --- it is hence consistent with our first argument. Note also that the innovations themselves, once adopted, tend to spread rapidly, leading again to a substantial convergence of forms. In absence of niches, the existing group barriers are indeed not effective to ``protect'' the diversity and the best canoe design rapidly spread from island to island fostered by the existing selection processes, e.g. differential fishing yields, migration success or simply survival of the sailors \citep[3418]{Rogers2008}.
The situation is markedly different for the symbolic traits of canoes, which presumably have little or no adaptive effect. In this case, a multitude of niches exists, all sharing the same level of adaptiveness (which is low, but this makes no difference as long as it is equal). This reflects in their large variability, in a condition where even weak barriers are sufficient to maintain significant group differences. Note also that symbolic markers are often used by humans in order to define group boundaries \citep[221-224]{RB2005}, a fact suggesting that these features are not only the product of different memes shared by different groups, but may represent barriers for the inter-group meme flow as well.
While this is just an example, it shows how our model can be tested using empirical data. The crucial point is to evaluate the existence of alternative niches for each trait included in the analysis. Once that a multi-niche environment is detected, the next question becomes whether a similar number of memetic groups exists. This crucially depends on the strength of boundaries among groups living in the environment under observation. The general prediction is that traits strictly associated with ``material'' aspects of live will vary only when real alternative niches exists, while symbolic traits are much more free to differentiate following the existing group boundaries.
\paragraph{Acknowledgements:}
The author gratefully acknowledges the comments received during the \emph{First International Workshop on Cultural Commons}, Torino, January 29--30, 2010. A special thanks goes to Lucia Tamburino for her comments and her help in developing part of the simulation codes.
\bibliographystyle{chicagoa}
|
\section{Introduction}
Let $N$ be a differential graded vector space and
let $M\subset N$ be a differential graded subspace such that the inclusion map
$\imath\colon M\to N$ is a quasi-isomorphism. The basic homology theory shows that there exists a homotopy $h\colon N\to N$ such that
$Id+dh+hd\colon N\to N$ is a projection onto $M$.
If $\tilde{d}$ is a new differential on $N$ such that $\partial=\tilde{d}-d$ is ``small'' in some appropriate sense, then the \emph{ordinary perturbation lemma} (Theorem~\ref{thm.pertlemma}) gives explicit functorial formulas, in terms of $\partial$ and $h$, for
a differential $\tilde{D}$ on $M$ and for an injective morphism of differential graded vector spaces
$\tilde{\imath}\colon (M,\tilde{D})\to (N,\tilde{d})$.\par
Has been pointed out by Huebschmann and Kadeishvili \cite{HK} that if $M,N$ are differential graded (co)algebra, and $h$ is a (co)algebra homotopy (Definition~\ref{def.coalgebracontraction}), then also
$\tilde{\imath}$ is a morphism of differential graded (co)algebras. This assumption are verified for instance when we consider the tensor coalgebras generated by $M,N$ and the natural extension of $h$ to $T(N)$ (this fact is referred as \emph{tensor trick} in the literature). Therefore the ordinary perturbation lemma can be easily used to prove Kadeishvili's theorem \cite{Kad80,Kad82} on the homotopy transfer of
$A_{\infty}$ structures (see also \cite{HK,hueainfty,KoSo,KonSoi,transfer,Merk}).\par
If we wants to use the same strategy for $L_{\infty}$-algebras, we have to face
the following problems:\begin{enumerate}
\item the tensor trick breaks down for symmetric
powers and coalgebra homotopies are not stable under symmetrization,
\item not every $L_{\infty}$-algebra is the symmetrization of an $A_{\infty}$-algebra.
\end{enumerate}
Therefore the proof of the homotopy transfer for $L_{\infty}$-algebras requires either a nontrivial additional work \cite{HS,pertlie,pertshlie} or a different approach, see e.g. \cite{fuka,K} and the arXiv version of \cite{cone}.
The aim of this paper is to show that the homotopy transfer for $L_{\infty}$-algebras (Theorem~\ref{thm.symmcoalgebrapert}) follows easily from a slight
modification (Theorem~\ref{thm.relapert}) of the ordinary perturbation lemma in which we assume that
$\tilde{d}$ is a differential when restricted to a differential graded subspace $A\subset N$ satisfying suitable properties.
The paper is written in a quite elementary style and we do not assume any knowledge of homological perturbation theory. We only assume that the reader is familiar with the basic properties of graded tensor and graded symmetric coalgebras. The bibliography contains the documents that have been more useful in the writing of this paper and it is necessarily incomplete; for more complete references the reader may consult
\cite{jimmurra,hueainfty}. I apologize in advance for every possible misattribution of previous results.
\bigskip
\section{The category of contractions}
Let $R$ be a fixed commutative ring; by a differential graded $R$-module
we mean a $\mathbb{Z}$-graded $R$-module $N=\oplus_{i\in \mathbb{Z}}N^i$ together a $R$-linear differential $d_N\colon N\to N$ of degree $+1$.
Given two differential graded $R$-modules
$M,N$ we denote by $\operatorname{Hom}_R^n(M,N)$ the module of $R$-linear maps of degree $n$:
\[ \operatorname{Hom}_R^n(M,N)=\{f\in\operatorname{Hom}_{R}(M,N)\mid f(M_i)\subset N_{i+n},\, \forall\; i\in\mathbb{Z}\}.\]
Notice that $\operatorname{Hom}_R^0(M,N)$ are the morphisms of graded $R$-modules and
\[\{f\in \operatorname{Hom}_R^0(M,N)\mid d_Nf=fd_M\}\]
is the set of cochain maps (morphisms of differential graded $R$-modules).
\begin{definition}[Eilenberg and Mac Lane {\cite[p. 81]{eilmactw}}]
A \emph{contraction}
is the data
\[ (\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi}, h)\]
where $M,N$ are differential graded $R$-modules, $h\in\operatorname{Hom}^{-1}_R(N,N)$ and $\imath,\pi$ are cochain maps such that:
\begin{enumerate}
\item (deformation retraction) $\;\pi\imath=\operatorname{Id}_{M}$,
$\;\imath\pi-\operatorname{Id}_{N}=d_{N}h+hd_{N}$,
\item (annihilation properties) $\;\pi h=h\imath=h^{2}=0$.
\end{enumerate}
\end{definition}
\begin{remark} In the original definition Eilenberg and Mac Lane do not require $h^2=0$; however, if $h$ satisfies the remaining 4 conditions, then $h'=hd_N h$ satisfies also the fifth (cf. \cite[Rem. 2.1]{pertshlie}).\end{remark}
\begin{definition}\label{def.morfismocontrazioni}
A \emph{morphism} of contractions
\[ f\colon
(\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi}, h)\to
(\xymatrix{A\ar@<.4ex>[r]^i&B\ar@<.4ex>[l]^p}, k)\]
is a morphism of differential graded $R$-modules $f\colon N\to B$ such that $fh=kf$.
Given a morphism of contractions as above we denote by $\hat{f}\colon M\to A$ the
morphism of differential graded $R$-modules $\hat{f}=pf\imath$.
\end{definition}
In the notation of Definition~\ref{def.morfismocontrazioni} it is easy to see that
the diagrams
\[ \xymatrix{M\ar[r]^{\hat{f}}\ar[d]^\imath&A\ar[d]^i\\
N\ar[r]^f&B}\qquad\qquad
\xymatrix{N\ar[r]^{f}\ar[d]^\pi&B\ar[d]^p\\
M\ar[r]^{\hat{f}}&A}
\]
are commutative. In fact
\[ i\hat{f}=ipf\imath=f\imath+(d_Bkf+kd_Bf)\imath=
f\imath+f(d_Nh+hd_N)\imath=f\imath+f(\imath\pi-\operatorname{Id}_N)\imath=
f\imath,\]
\[ \hat{f}\pi=pf\imath\pi=pf(\operatorname{Id}_N+d_Nh+hd_N)=
pf+p(d_Bk+kd_B)f=pf+p(ip-\operatorname{Id}_B)=pf.\]
\begin{definition}\label{def.composizionecontrazioni}
The \emph{composition} of contractions is defined as
\[
(\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi}, h)\circ
(\xymatrix{N\ar@<.4ex>[r]^i&P\ar@<.4ex>[l]^p}, k)=
(\xymatrix{M\ar@<.4ex>[r]^{i\imath}&P\ar@<.4ex>[l]^{\pi p}}, k+ihp)
\]
\end{definition}
\bigskip
Given two contractions
$(\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi}, h)$ and
$(\xymatrix{A\ar@<.4ex>[r]^i&B\ar@<.4ex>[l]^p}, k)$ we define
their tensor product as
\[ (\xymatrix{M\otimes_RA\ar@<.4ex>[r]^{\imath\otimes i}&
N\otimes_RB\ar@<.4ex>[l]^{\pi\otimes p}}, h\ast k),
\qquad h\ast k=\imath\pi\otimes k+h\otimes \operatorname{Id}_B.\]%
It is straightforward to verify that the tensor product of two contractions is a contraction, it is bifunctorial and, up to the canonical isomorphism $(L\otimes_R M)\otimes_R N\cong L\otimes_R (M\otimes_R N)$, it is associative.\par
Given a contraction $(\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi}, h)$, its tensor $n$th power is
\[ \tensor^n_R(\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi}, h)=
(\xymatrix{M^{\otimes n}\ar@<.4ex>[r]^{\imath^{\otimes n}}&N^{\otimes n}
\ar@<.4ex>[l]^{\pi^{\otimes n}}}, T^nh),\]
where
\[ T^nh=\sum_{i=1}^{n}(\imath\pi)^{\otimes i-1}
\otimes h\otimes \operatorname{Id}_N^{\otimes n-i}.\]
The tensor product allows to define naturally the notion of algebra and coalgebra contraction; we consider here only the case of coalgebras.
\begin{definition}\label{def.coalgebracontraction}
Let $N$ be a differential graded coalgebra over a commutative ring $R$ with coproduct $\Delta\colon N\to N\otimes_R N$.
We shall say that a contraction $(\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi}, h)$
is a \emph{coalgebra contraction} if
\[ \Delta \colon (\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi}, h)\to
(\xymatrix{M\otimes_RM\ar@<.4ex>[r]^{\imath\otimes \imath}&
N\otimes_RN\ar@<.4ex>[l]^{\pi\otimes \pi}}, h\ast h)\]
is a morphism of contractions.
\end{definition}
Notice that if $\Delta$ is a morphism of contractions then $\hat{\Delta}$ is a coproduct and
$\pi,\imath$ are morphisms of differential graded
coalgebras. Conversely,
a contraction $(\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi}, h)$ is a coalgebra contraction if $\pi,\imath$ are morphisms of differential graded
coalgebras and
\[ (\imath\pi\otimes h+h\otimes\operatorname{Id}_N)\circ\Delta=
\Delta\circ h.\]%
\begin{example}[tensor trick]\label{exa.tensorcoalgebracontractions}
Given a contraction $(\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi}, h)$ of differential graded $R$-modules, we can
consider the \emph{reduced tensor coalgebra}
\[\bar{T}(N)=\somdir_{n=1}^{\infty}\tensor^n_RN\] with the
coproduct
\[ \mathfrak{a}(x_1\otimes\cdots\otimes x_n)=\sum_{i=1}^{n-1}
(x_1\otimes\cdots\otimes x_i)\otimes (x_{i+1}\otimes\cdots\otimes x_n).\]
We have seen that there exists a contraction
\[(\xymatrix{\bar{T}(M)\ar@<.4ex>[r]^{T(\imath)}&\bar{T}(N)
\ar@<.4ex>[l]^{T(\pi)}}, Th)\;,\]
where $T(\imath)=\sum \imath^{\otimes n}$, $T(\pi)=\sum \pi^{\otimes n}$ and
$Th=\sum_n T^nh$.\par
We want to prove that $(\xymatrix{\bar{T}(M)\ar@<.4ex>[r]^{T(\imath)}&\bar{T}(N)
\ar@<.4ex>[l]^{T(\pi)}}, Th)$ is a coalgebra contraction, i.e. that
\[ (T(\imath\pi)\otimes Th+Th\otimes\operatorname{Id})\circ\mathfrak{a}=
\mathfrak{a}\circ Th.\]
Let $n$ be a fixed positive integer, writing
\[ T^nh=\sum_{i=1}^{n}T^n_ih\;,\qquad
T^n_ih=(\imath\pi)^{\otimes i-1}\otimes h\otimes \operatorname{Id}_{N}^{\otimes n-i},\]
for every $i=1,\ldots,n$ we have
\[ \mathfrak{a} \circ T^n_ih=\sum_{j=1}^{i-1}(\imath\pi)^{\otimes j}\otimes T^{n-j}_{i-j}h+
\sum_{j=i}^{n-1} T^{j}_{i}h\otimes \operatorname{Id}_N^{\otimes n-j}.\]
Therefore
\begin{align*}\mathfrak{a} \circ T^nh=
\sum_{i=1}^n\mathfrak{a} \circ T^n_ih
&=\sum_{i=1}^n\sum_{j=1}^{i-1}(\imath\pi)^{\otimes j}\otimes T^{n-j}_{i-j}h+
\sum_{i=1}^n\sum_{j=i}^{n-1} T^{j}_{i}h\otimes \operatorname{Id}_N^{\otimes n-j}\\
&=\sum_{j=1}^{n-1}\sum_{i=j+1}^{n}(\imath\pi)^{\otimes j}\otimes T^{n-j}_{i-j}h+
\sum_{j=1}^{n-1}\sum_{i=1}^{j} T^{j}_{i}h\otimes \operatorname{Id}_N^{\otimes n-j}\\
&=\sum_{j=1}^{n-1}(\imath\pi)^{\otimes j}\otimes (\sum_{i=1}^{n-j} T^{n-j}_{i}h)+
\sum_{j=1}^{n-1}(\sum_{i=1}^{j} T^{j}_{i}h)\otimes \operatorname{Id}_N^{\otimes n-j}\\
&=\sum_{j=1}^{n-1}(\imath\pi)^{\otimes j}\otimes T^{n-j}h+
\sum_{j=1}^{n-1}T^{j}h\otimes \operatorname{Id}_N^{\otimes n-j}.
\end{align*}
It is now sufficient to sum over $n$.\end{example}
\bigskip
\section{Review of ordinary homological perturbation theory}
\textbf{Convention:} \emph{In order to simplify the notation, from now on, and unless otherwise stated, for every
contraction $(\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi}, h)$ we assume that
$M$ is a submodule of $N$ and $\imath$ the inclusion.}
\bigskip\par
Given a contraction $(\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi}, h)$ of differential graded $R$-modules and a morphism $\partial\in \operatorname{Hom}^1_R(N,N)$, the ordinary homological perturbation theory consists is a series of statements about the maps
\begin{equation}\label{equ.inclusioneperturbata}
\imath_{\partial}=\sum_{n\ge 0}(h\partial)^n\imath\;\in \operatorname{Hom}^0_R(M,N),
\end{equation}
\begin{equation}
\qquad\pi_{\partial}=\sum_{n\ge 0}\pi(\partial h)^n\;\in \operatorname{Hom}^0_R(N,M),
\end{equation}
\begin{equation}
D_{\partial}=\pi\partial\imath_{\partial}=\pi_{\partial}\partial\imath\;\in \operatorname{Hom}^1_R(M,M),
\end{equation}
In order to have the above maps defined we need to impose some extra assumption. This may done by considering filtered contractions of complete modules (as in \cite{HK}) or
by imposing a sort of local nilpotency for the operators $h\partial, \partial h$.
\begin{definition}\label{def.mezzoperturbation}
Given a contraction $(\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi},h)$ denote
\[ \mathcal{N}(N,h)=\{\partial\in \operatorname{Hom}^1_R(N,N)\mid \cup_n\ker((h\partial)^n\imath)
=M,\; \cup_n\ker(\pi(\partial h)^n)=N\}.\]
\end{definition}
It is plain that the maps $\imath_{\partial}, \pi_{\partial}$ and $D_{\partial}$ are well defined for every $\partial\in\mathcal{N}(N,h)$. Moreover they are functorial in the following sense: given a morphism of contractions
\[ f\colon (\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi},h)
\to (\xymatrix{A\ar@<.4ex>[r]^i&B\ar@<.4ex>[l]^p},k)\]
and two elements $\partial\in \mathcal{N}(N,h)$, $\delta\in \mathcal{N}(B,k)$ such that
$f\partial=\delta f$ we have
\[f\imath_{\partial}=\sum_{n\ge 0}f(h\partial)^n\imath=\sum_{n\ge 0}(k\delta)^nf\imath=
\sum_{n\ge 0}(k\delta)^ni\hat{f}=i_{\delta}\hat{f}.\]
Similarly we have
$\hat{f}\pi_{\partial}=p_{\delta}f$,
$\hat{f}D_{\partial}=D_{\delta}\hat{f}$.
\begin{lemma}\label{lem.injectivity} Let $(\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi},h)$ be a contraction and $\partial\in \mathcal{N}(N,h)$. Then $\imath_{\partial}$ is injective and
\[ \pi_{\partial}\imath_{\partial}=\pi\imath=\operatorname{Id}_M.\]
\end{lemma}
\begin{proof} Immediate consequence of annihilation properties.
It is useful to point out that the proof of the injectivity of $\imath_{\partial}$ does not depend on the annihilation properties.
Assume $\imath_{\partial}(x)=0$ and let $s\ge 0$ be the
minimum integer such that $(h\partial)^s\imath(x)=0$. If $s>0$ then
\[ 0=(h\partial)^{s-1}\imath_{\partial}(x)=(h\partial)^{s-1}\imath(x)+
\sum_{k\ge s}(h\partial)^{k}\imath(x)= (h\partial)^{s-1}\imath(x)\]
giving a contradiction. Hence $s=0$ and $\imath(x)=0$.
\end{proof}
\begin{proposition}\label{prop.compositioncompatibility}
The formula \ref{equ.inclusioneperturbata} is compatible with composition of contractions. More precisely, if
\[ (\xymatrix{L\ar@<.4ex>[r]^i&M\ar@<.4ex>[l]^p}, k)\circ
(\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi}, h)
=
(\xymatrix{L\ar@<.4ex>[r]^{\imath i}&N\ar@<.4ex>[l]^{p\pi }}, h+\imath k \pi)
\]
then $(\imath i)_{\partial}=\imath_{\partial} i_{D_{\partial}}$,
provided that all terms of the equation are defined.
\end{proposition}
\begin{proof} We have
\begin{align*}
\imath_{\partial} i_{D_{\partial}}=&\sum_{n\ge 0}(h\partial)^n\imath
\sum_{m\ge 0}(kD_{\partial})^m i\\
&=\sum_{n\ge 0}(h\partial)^n
\sum_{m\ge 0}\imath(k\pi\partial \sum_{s\ge 0}(h\partial)^s\imath)^m i\\
&=\sum_{n\ge 0}(h\partial)^n
\sum_{m\ge 0}(\imath k\pi\partial \sum_{s\ge 0}(h\partial)^s)^m\imath i\\
&=\sum_{n\ge 0}(h\partial+\imath k\pi\partial )^n\imath i\\
&=(\imath i)_{\partial}.
\end{align*}
\end{proof}
\begin{proposition}\label{prop.huebKad}
Let $(\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi},h)$ be a coalgebra contraction
and $\partial\in \mathcal{N}(N,h)$. If $\partial$ is a coderivation then
$\imath_{\partial}$ and $\pi_{\partial}$ are morphisms of graded coalgebras and $D_{\partial}$ is a coderivation.
\end{proposition}
\begin{proof} Consider the contraction
\[(\xymatrix{M\otimes_RM\ar@<.4ex>[r]^{\imath\otimes\imath}&
N\otimes_RN\ar@<.4ex>[l]^{\pi\otimes\pi}},k), \quad\text{ where }\quad
k=h*h=\imath\pi\otimes h+h\otimes \operatorname{Id}_N,\]
and $\delta=\partial\otimes\operatorname{Id}_N+\operatorname{Id}_N\otimes\partial$. In order to prove that $\delta\in \mathcal{N}(N\otimes_RN,k)$ we show that for every integer $n\ge 0$ we have
\[ (k\delta)^n(\imath\otimes\imath)=
\sum_{i+j=n}(h\partial)^i\imath\otimes (h\partial )^j\imath\;,\qquad
(\pi\otimes\pi)(\delta k)^n=
\sum_{i+j=n}\pi(\partial h)^i\otimes\pi (\partial h)^j\;.
\]
We prove here only the first equality by induction on $n$; the second is completely similar and left to the reader. Since
\[ k\delta=h\partial\otimes\operatorname{Id}_N+h\otimes\partial-\imath\pi\partial\otimes
h+\imath\pi\otimes h\partial,\]%
according to annihilation properties we have:
\[h\partial\otimes\operatorname{Id}_N\left(\sum_{i+j=n}(h\partial )^i\imath\otimes(h\partial
)^j\imath\right)=\sum_{i+j=n}(h\partial )^{i+1}\imath\otimes(h\partial
)^j\imath,\]
\[h\otimes\partial\left(\sum_{i+j=n}(h\partial )^i\imath\otimes(h\partial )^j\imath\right)
=0,\qquad
\imath\pi\partial\otimes h\left(\sum_{i+j=n}(h\partial )^i\imath\otimes(h\partial
)^j\imath\right)=0,\]
\[\imath\pi\otimes h\partial\left(\sum_{i+j=n}(h\partial )^i\imath\otimes(h\partial
)^j\imath\right)=\imath\otimes (h\partial )^{n+1}\imath.\]
Therefore
\[ (\imath\otimes\imath)_{\delta}=
\sum_{n\ge 0}(k\delta)^n(\imath\otimes\imath)=
\sum_{i,j\ge 0}(h\partial)^i\imath\otimes (h\partial )^j\imath=
\imath_{\partial}\otimes\imath_{\partial}\;,
\]
\[ (\pi\otimes\pi)_{\delta}=
\sum_{n\ge 0}(\pi\otimes\pi)(\delta k)^n=
\sum_{i,j\ge 0}\pi(\partial h)^i\otimes \pi(\partial h)^j=
\pi_{\partial}\otimes\pi_{\partial}\;.
\]
Denoting by $\Delta\colon N\to N\otimes_R N$ the coproduct,
since $\partial$ is a coderivation we have $\delta\Delta=\Delta\partial$; since $\Delta$ is a morphism of contractions we have by functoriality
\[ \Delta\imath_{\partial}=(\imath\otimes\imath)_{\delta}\hat{\Delta}=
(\imath_{\partial}\otimes\imath_{\partial})\hat{\Delta},\qquad \hat{\Delta}\pi_{\partial}=(\pi\otimes\pi)_{\delta}\Delta=
(\pi_{\partial}\otimes\pi_{\partial})\Delta,\]
and then $\imath_{\partial}, \pi_{\partial}$ are morphisms of coalgebras.
Finally $D_{\partial}$ is a coderivation because it is the composition of the coderivation $\partial$ and the two morphisms of coalgebras $\imath_{\partial}$ and $\pi$.
\end{proof}
A proof of Proposition~\ref{prop.huebKad} is given in \cite{HK} under the unnecessary assumption that
$(d+\partial)^2=0$.
\begin{definition} Let $N$ be a differential graded $R$-module.
A \emph{perturbation} of the differential $d_N$
is a linear map $\partial\in \operatorname{Hom}^1_R(N,N)$ such that
$(d_N+\partial)^2=0$.
\end{definition}
\begin{theorem}[Ordinary perturbation lemma]\label{thm.pertlemma}
Let
$(\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi},h)$ be a contraction and let $\partial\in
\mathcal{N}(N,h)$ be a perturbation of the differential $d_N$.
Then $D_{\partial}$ is a perturbation of
$d_{M}=\pi d_N\imath$ and
\[ \pi_{\partial}\colon (N,d_N+\partial)\to (M,d_M+D_{\partial}),\qquad
\imath_{\partial}\colon (M,d_M+D_{\partial})\to (N,d_N+\partial)\]%
are morphisms of differential graded $R$-modules.
\end{theorem}
\begin{proof} See \cite{HK,jimmurra} and references therein for proofs and history. We prove again this result in Remark~\ref{rem.cpl} as a particular case of the relative perturbation lemma.
\end{proof}
\begin{remark} If $\cup_n\ker(h\partial)^n=N$, and $\partial$ is a perturbation of $d_N$,
then $\imath_{\partial}$ is the unique morphism of graded $R$-modules
$M\to N$ whose image is a subcomplex of $(N,d_N+\partial)$ and satisfying the ``gauge fixing'' condition
\[ h\imath_{\partial}=0,\qquad \pi\imath_{\partial}=\operatorname{Id}_M.\]
In fact $h(d_N+\partial)\imath_{\partial}=0$
and then
\begin{align*}
\imath_{\partial}=&\imath_{\partial}+hd_N\imath_{\partial}+h\partial\imath_{\partial}=(\imath\pi-d_Nh)\imath_{\partial}+
h\partial\imath_{\partial}=\imath+(h\partial)\imath_{\partial}\\
=&(\operatorname{Id}_N-h\partial)^{-1}\imath.
\end{align*}
Similarly $\pi_{\partial}$ is the unique morphism of graded $R$-modules
$M\to N$ whose kernel is a subcomplex of $(N,d_N+\partial)$ and satisfying
\[ \pi_{\partial}h=0,\qquad \pi_{\partial}\imath=\operatorname{Id}_M.\]
\end{remark}
The coalgebra perturbation lemma cited in the abstract is obtained by putting together
Proposition~\ref{prop.huebKad} and Theorem~\ref{thm.pertlemma}.
\medskip
\section{The relative perturbation lemma}
\label{sec.relativepertlemma}
\begin{definition}\label{def.relativeperturbation}
Let $N$ be a differential graded $R$-module and $A\subset N$ a
differential graded submodule. A morphism $\partial\in \operatorname{Hom}^1_R(N,N)$ is called a \emph{perturbation of $d_N$ over $A$} if
\[ \partial
(A)\subset A\quad\text{and}\quad (d_N+\partial)^2(A)=0.\]
\end{definition}
\begin{remark} The meaning of Definition~\ref{def.relativeperturbation} becomes more clear when we impose some extra assumption on $\partial$. For instance, if $N$ is a differential graded coalgebra and $\partial$ is a coderivation, then in general does not exist any coderivation $\delta$ of $N$ such that $\delta_{|A}=\partial_{|A}$ and
$(d_N+\delta)^2=0$. An explicit example of this phenomenon will be described in
Section~\ref{sec.coderivation}.\end{remark}
\begin{theorem}[Relative perturbation lemma]\label{thm.relapert}
Let $(\xymatrix{M\ar@<.4ex>[r]^\imath&N\ar@<.4ex>[l]^\pi},h)$ be a contraction with $M\subset N$ and $\imath$ the inclusion. Let $A\subset N$ be a differential graded submodule
and $\partial\in \mathcal{N}(N,h)$ a perturbation of $d_N$
over $A$. Assume moreover that:
\begin{enumerate}
\item $\pi(A)\subset A\cap M$.
\item $\imath_{\partial}(A\cap M)\subset A$.
\end{enumerate}
Then
\[ D_{\partial}=\sum_{n\ge 0}\pi\partial (h\partial)^n\imath=\sum_{n\ge 0}\pi(\partial
h)^n\partial\imath\in \operatorname{Hom}^1_R(M,M),\]
is a perturbation of $d_M$ over $A\cap M$ and
\[ \imath_{\partial}=\sum_{n\ge 0}(h\partial)^n\imath\colon (A\cap M,d_M+D_{\partial})\to (A,d_N+\partial)\]%
is a morphisms of differential graded $R$-modules.
\end{theorem}
\begin{remark} It is important to point out that
we do not require that $h(A)\subset A$ but only the weaker assumption
$\imath_{\partial}(M\cap A)\subset A$.\end{remark}
\begin{proof} We first note that
$D_{\partial}=\pi\partial\imath_{\partial}$ and then
$D_{\partial}(A\cap M)\subset A\cap M$. In order to simplify the notation we denote $d=d_N$ and $I=Id_N$.
Setting $\psi=\partial^2+d\partial+\partial d\in \operatorname{Hom}^2_R(N,N)$ we have the formula
\begin{equation}\label{equ.l4}
\sum_{n,m\ge 0}(\partial h)^n\partial\imath\pi\partial(h\partial)^m=
\sum_{n,m\ge 0}(\partial h)^n\psi(h\partial)^m
-\sum_{m\ge 0} d\partial(h\partial)^m-\sum_{n\ge 0}(\partial h)^n\partial d.
\end{equation}
In fact, since $\imath\pi=I+hd+dh$, we have
\[ \partial\imath\pi\partial=\partial(I+hd+dh)\partial=\partial^2+\partial hd\partial+\partial dh\partial=
\psi-(I-\partial h)d\partial-\partial d(I-h\partial)\]
and therefore
\begin{align*}
\sum_{n,m\ge 0}&(\partial h)^n\partial\imath\pi\partial(h\partial)^m\\
=&\sum_{n,m\ge 0}(\partial
h)^n\psi(h\partial)^m-\sum_{n,m\ge 0}(\partial h)^n(I-\partial h)d\partial(h\partial)^m
-\sum_{n,m\ge 0}(\partial h)^n\partial d(I-h\partial)(h\partial)^m\\
=&\sum_{n,m\ge 0}(\partial
h)^n\psi(h\partial)^m-\sum_{m\ge 0}d\partial(h\partial)^m
-\sum_{n\ge 0}(\partial h)^n\partial d\;.
\end{align*}
We have
\begin{align*}
(d+\partial)\imath_{\partial}=&\sum_{m\ge
0}d(h\partial)^m\imath+\sum_{m\ge 0}\partial(h\partial)^m\imath
=d\imath+\sum_{m\ge
0}dh\partial(h\partial)^m\imath+\sum_{m\ge 0}\partial(h\partial)^m\imath\\
=&d\imath +\sum_{m\ge 0}(I+dh)\partial(h\partial)^m\imath
=d\imath+\sum_{m\ge 0}(\imath\pi-hd)\partial(h\partial)^m\imath\;,
\end{align*}
\begin{align*}
\imath_{\partial}(d_M+D_{\partial})\!=&
\sum_{n\ge 0}(h\partial)^n\imath d_M+\sum_{n,m\ge 0}(h\partial)^n\imath\pi\partial(h\partial)^m\imath\\
=&\sum_{n\ge 0}(h\partial)^n\imath d_M+\!
\sum_{m\ge 0}\imath\pi\partial(h\partial)^m\imath+
h\sum_{n,m\ge 0}(\partial h)^n\partial\imath\pi\partial(h\partial)^m\imath\qquad\qquad\\
=&\sum_{n\ge 0}(h\partial)^nd\imath+\!\sum_{m\ge 0}\imath\pi\partial(h\partial)^m\imath
+\!\sum_{n\ge 0}h(\partial h)^n\psi\imath_{\partial}
-\!\sum_{m\ge 0} hd\partial(h\partial)^m\imath-\!\sum_{n\ge 0}h(\partial h)^n\partial d\imath\\
=&d\imath+\sum_{m\ge 0}(\imath\pi-hd)\partial(h\partial)^m\imath
+\sum_{n\ge 0}h(\partial h)^n\psi\imath_{\partial},
\end{align*}
and therefore
\[ \imath_{\partial}(d_M+D_{\partial})-(d+\partial)\imath_{\partial}=\sum_{n\ge 0}h(\partial h)^n\psi\imath_{\partial}.\]
In particular,
for every $x\in M\cap A$ we have $\psi\imath_{\partial}(x)=0$ and then
\[ \imath_{\partial}(d_M+D_{\partial})(x)=(d+\partial)\imath_{\partial}(x).\]
Now we prove that $D_{\partial}$ is perturbation of $d_M$ over $M\cap A$, i.e. that
$(d_M+D_{\partial})^2x=0$
for every $x\in M\cap A$. Since $\pi h=0$ we have $\pi\imath_{\partial}=\pi\imath$ and then
$\imath_{\partial}$ is injective.
If $x\in M\cap A$ we have
\[ \imath_{\partial}(d_M+D_{\partial})^2x=(d+\partial)\imath_{\partial}(d_M+D_{\partial})x=
(d+\partial)^2\imath_{\partial}x=0.\]
\end{proof}
\begin{remark}\label{rem.cpl} In the set-up of Theorem~\ref{thm.relapert}, if $h(A)\subset A$ then also $\pi_{\partial}\colon (A,d+\partial)\to (A\cap M, d_M+D_{\partial})$ is a morphism of differential graded $R$-modules. In fact, under this additional assumption we have
\[ \pi_{\partial}(A)=\sum_{n\ge0} \pi(\partial h)^n(A)\subset A\cap M,\qquad
\sum_{n,m\ge 0}(\partial h)^n\psi(h\partial)^m h(A)=0,\]
and therefore in $A$ the following equalities hold:
\begin{multline*}
\pi_{\partial}(d+\partial)=\sum_{n\ge 0}\pi (\partial h)^n d+\sum_{n\ge 0}\pi
(\partial h)^n\partial=\pi d+\sum_{n\ge 0}\pi (\partial h)^n\partial hd+\sum_{n\ge
0}\pi (\partial h)^n\partial\\
=\pi d+\sum_{n\ge 0}\pi (\partial h)^n\partial(I+hd)= \pi d+\sum_{n\ge 0}\pi
(\partial h)^n\partial(\imath\pi-dh).
\end{multline*}
\begin{align*}
(d+D_{\partial})\pi_{\partial}=&\sum_{n\ge 0}\pi d(\partial h)^n+ \sum_{n,m\ge 0}\pi(\partial
h)^n\partial\imath\pi(\partial h)^m\\
=&\sum_{n\ge 0}\pi d(\partial h)^n+ \sum_{n\ge 0}\pi(\partial
h)^n\partial\imath\pi +\sum_{n\ge 0,m\ge 1}\pi(\partial
h)^n\partial\imath\pi(\partial h)^m\\
=&\sum_{n\ge 0}\pi d(\partial h)^n+ \sum_{n\ge 0}\pi(\partial
h)^n\partial\imath\pi +\sum_{n,m\ge 0}\pi(\partial
h)^n\partial\imath\pi\partial (h\partial)^mh\\
=&\sum_{n\ge 0}\pi d(\partial h)^n+ \sum_{n\ge 0}\pi(\partial
h)^n\partial\imath\pi - \sum_{m\ge 0}\pi d\partial(h\partial)^mh
-\sum_{n\ge 0}\pi (\partial h)^n\partial dh\\
=&\left(\sum_{n\ge 0}\pi d(\partial h)^n- \sum_{m\ge 0}\pi
d\partial(h\partial)^mh\right)+\sum_{n\ge 0}\pi(\partial h)^n\partial\imath\pi
-\sum_{n\ge 0}\pi (\partial h)^n\partial dh\\
=& \pi d+\sum_{n\ge 0}\pi (\partial h)^n\partial(\imath\pi-dh).
\end{align*}
\end{remark}
\begin{remark} It is straightforward to verify that all the previous proofs also work for
the weaker notion of contraction where the condition $\pi\imath=Id_M$ is replaced with
\emph{$\imath$ is injective and $\imath(M)$ is a direct summand of $N$ as graded $R$-module.}
\end{remark}
\bigskip
\section{Review of reduced symmetric coalgebras and their coderivations}
\label{sec.coderivation}
From now on we assume that $R=\mathbb{K}$ is a field of characteristic 0. Given a graded vector space
$V$, the \emph{twist map}
\[{\pmb{\mathtt{tw}}}\colon V\otimes V\to V\otimes V,\qquad
{\pmb{\mathtt{tw}}}(v\otimes w)=(-1)^{\deg(v)\deg(w)}w\otimes v,\]
extends naturally to an action of the symmetric group $\Sigma_n$ on the tensor product
$\bigotimes^nV$:
\[ \sigma_{{\pmb{\mathtt{tw}}}}(v_{1}\otimes\cdots\otimes v_{n})=
\pm
\;v_{\sigma^{-1}(1)}\otimes\cdots\otimes v_{\sigma^{-1}(n)},\qquad \sigma\in \Sigma_n.\]
We will denote by $\bigodot^nV=(\bigotimes^nV)^{\Sigma_n}$ the subspace of invariant tensors. Notice that if $W\subset V$ is a graded subspace, then $\bigodot^nW=\bigodot^nV\cap \bigotimes^nW$.
It is easy to see that the subspace
\[ \bar{S}(V)=\somdir_{n=1}^{\infty}\symmetric^nV\subset
\somdir_{n=1}^{\infty}\tensor^nV=\bar{T}(V)\]
is a graded subcoalgebra, called the \emph{reduced symmetric coalgebra} generated by $V$.
Let's denote by $p\colon \bar{T}(V)\to V$ the projection; we will also denote by
$p\colon \bar{S}(V)\to V$ the restriction of the projection to symmetric tensors. The following well known properties hold (for proofs see e.g. \cite{defomanifolds}):
\begin{enumerate}
\item Given a morphism of graded coalgebras $F\colon \bar{T}(V)\to \bar{T}(W)$ we have
$F(\bar{S}(V))\subset \bar{S}(W)$.
\item Given a morphism of graded vector spaces $f\colon \bar{T}(V)\to W$
there exists an unique morphism of graded coalgebras $F\colon \bar{T}(V)\to \bar{T}(W)$ such that $f=pF$.
\item Given a morphism of graded vector spaces $f\colon \bar{S}(V)\to W$
there exists an unique morphism of graded coalgebras $F\colon \bar{S}(V)\to \bar{S}(W)$ such that $f=pF$.
\end{enumerate}
Similar results hold for coderivations. More precisely for
every map $q\in\operatorname{Hom}^k(\bar{T}(V),V)$ there exists an unique coderivation
$Q\colon \bar{T}(V)\to \bar{T}(V)$ of degree $k$ such that $q=pQ$. The coderivation $Q$ is given by the explicit formula
\begin{equation}\label{equ.coder}
Q(a_{1}\otimes\cdots\otimes a_{n})
=\sum_{l=1}^{n}\sum_{i=0}^{n-l}(-1)^{k(\bar{a_{1}}+\cdots+\bar{a_{i}})}
a_{1}\otimes\cdots\otimes a_{i}\otimes
q(a_{i+1}\otimes\cdots\otimes a_{i+l})\otimes\cdots\otimes a_{n}\;,
\end{equation}
where $\bar{a_i}=\deg(a_i)$. Moreover $Q(\bar{S}(V))\subset \bar{S}(V)$ and the restriction
of $Q$ to $\bar{S}(V)$ depends only on the restriction of $q$ on $\bar{S}(V)$. In particular every coderivation of $\bar{S}(V)$ extends to a coderivation of $\bar{T}(V)$.
\begin{definition} A coderivation $Q$ of degree $+1$ is called a \emph{codifferential} if
$Q^2=0$.\end{definition}
\begin{lemma} A coderivation $Q$ of degree $+1$ is a codifferential if and only if
$pQ^2=0$.\end{lemma}
\begin{proof} The space of coderivations of a graded coalgebra is closed under the bracket
\[ [Q,R]=QR-(-1)^{\deg(Q)\deg(R)}RQ\]
and therefore if $Q$ is a coderivation of odd degree, then its square $Q^2=[Q,Q]/2$ is again a coderivation.\end{proof}
Every codifferential on $\bar{T}(V)$ induces by restriction a codifferential on
$\bar{S}(V)$. Conversely it is generally false that a codifferential on
$\bar{S}(V)$ extends to a codifferential on $\bar{T}(V)$. This is well known to experts; however we will give here
an example of this phenomenon for the lack of suitable references.\par
We restrict our attention to graded vector spaces concentrated in degree $-1$, more precisely we assume that
$V=L[1]$, where $L$ is a vector space and $[1]$ denotes the shifting of the degree, i.e.
$L[1]^{i}=L^{i+1}$.
Under this assumption every codifferential in $\bar{T}(V)$ (resp.:
$\bar{S}(V)$) is determined by a linear map $q\colon \bigotimes^2 V\to V$ (resp.:
$q\colon \bigodot^2 V\to V$) of degree $+1$.
\begin{lemma} In the above assumption:
\begin{enumerate}
\item The map
\[L\times L\to L,\qquad xy=q(x\otimes y),\]
is an associative product if and only if $q$ induces a codifferential in $\bar{T}(V)$.
\item The map
\[L\times L\to L,\qquad [x,y]=q(x\otimes y-y\otimes x)=xy-yx,\]
is a Lie bracket if and only if $q$ induces a codifferential in $\bar{S}(V)$.
\end{enumerate}
\end{lemma}
\begin{proof} We have seen that $Q$ is a codifferential in $\bar{T}(V)$ if and only if
$pQ^2=qQ\colon \tensor^3V\to V$ is the trivial map. It is sufficient to observe that
\[qQ(x\otimes y\otimes z)=q(q(x\otimes y)\otimes z)-q(x\otimes q(y\otimes z))=
(xy)z-x(yz).\]
Similarly $Q$ is a codifferential in $\bar{S}(V)$ if and only if for every $x_1,x_2,x_3$ we have
\begin{align*}
0=&qQ\left(\sum_{\sigma\in \Sigma_3}(-1)^{\sigma}x_{\sigma(1)}\otimes x_{\sigma(2)}\otimes x_{\sigma(3)}\right)\\
=& \sum_{\sigma\in \Sigma_3}(-1)^{\sigma} ((x_{\sigma(1)}x_{\sigma(2)})x_{\sigma(3)}-
x_{\sigma(1)}(x_{\sigma(2)}x_{\sigma(3)}))\\
=&[[x_1,x_2],x_3]+[[x_2,x_3],x_1]+[[x_3,x_1],x_2]
\end{align*}
\end{proof}
Therefore every Lie bracket on $L$ not induced by an associative product gives a codifferential on $\bar{S}(L[1])$ which does not extend to a codifferential on
$\bar{T}(L[1])$.
\begin{example} Let $\mathbb{K}$ be a field of characteristic $\not=2$ and $L$ a vector space of dimension $3$ over $\mathbb{K}$ with basis $A,B,H$.
Then does not exist any associative product on $L$ such that
\[ AB-BA=H,\qquad HA-AH=2A,\qquad HB-BH=-2B.\]
We prove this fact by contradiction: assume that there exists an associative product as above,
then the pair $(L,[,])$, where $[X,Y]=XY-YX$, is a Lie algebra isomorphic to $sl_2(\mathbb{K})$. Writing
\[ H^2=\gamma_1 A+\gamma_2 B+\gamma H\]
we have
\[ 0=[H^2,H]=\gamma_1 [A,H]+\gamma_2 [B,H]\]
and therefore $\gamma_1=\gamma_2=0$, $H^2=\gamma H$.
Possibly acting with the Lie automorphism
\[ A\mapsto B,\qquad B\mapsto A,\qquad H\mapsto -H,\]
it is not restrictive to assume $\gamma\neq -1$.
Since $[AH,H]=[A,H]H=-2AH$, writing $AH=xA+yB+zH$ for some $x,y,z\in\mathbb{K}$ we have
\[ 0=[AH,H]+2AH=x[A,H]+y[B,H]+2xA+2yB+2zH=4yB+2zH\]
giving $y=z=0$ and $AH=xA$. Moreover $2A^2=A[H,A]=[AH,A]=[xA,A]=0$ and then $A^2=0$.
Since
\[ 0=A(H^2)-(AH)H=\gamma AH-x AH=(\gamma x-x^2)A\]
we have either $x=0$ or $x=\gamma$. In both cases $x\not=-1$ and then $AH+HA=(2x+2)A\not=0$.
This gives a contradiction since
\[ -AH=A(AB-H)=ABA=(BA+H)A=HA.\]
\end{example}
\bigskip
\section{The $L_{\infty}$-algebra perturbation lemma}
The bar construction gives an equivalence from the category of $L_{\infty}$-algebras
and the category of differential graded reduced symmetric coalgebras (see e.g. \cite{cone,fuka,K}).\par
According to Formula~\ref{equ.coder}, every coderivation $Q\colon \bar{T}(V)\to \bar{T}(V)$ of degree $+1$ can be uniquely decomposed as
$Q=d+\partial$, where
\[d(\tensor^n V)\subset \tensor^n V,\qquad
\partial(\tensor^n V)\subset \somdir_{i=1}^{n-1}\tensor^i V,\qquad \forall\; n>0.\]
and
\[d(a_{1}\otimes\cdots\otimes a_{n})
=\sum_{i=0}^{n-1}(-1)^{\bar{a_{1}}+\cdots+\bar{a_{i}}}
a_{1}\otimes\cdots\otimes a_{i}\otimes
d_1(a_{i+1})\otimes a_{i+2}\otimes\cdots\otimes a_{n}\]
where $d_1=Q_{|V}\colon V\to V$.
If $Q$ is a codifferential on $\bar{T}(V)$ then $d^2(V)=0$, $d$ is the natural differential on the tensor powers of the complex $(V,d_1)$ and $\partial$ is a perturbation of $d$.
If $Q$ is a codifferential on $\bar{S}(V)$ then $d^2(V)=0$ and therefore $d$ is the natural differential on the symmetric powers of the complex $(V,d_1)$ and $\partial$ is a perturbation of $d$ over $\bar{S}(V)$.
\begin{theorem}\label{thm.symmcoalgebrapert}
In the above notation, let $Q=d+\partial$ be a coderivation of degree +1 on
$\bar{T}(V)$ which is a codifferential on $\bar{S}(V)$. Let $W$ be a differential
graded subspace of $(V,d)$ and let
$(\xymatrix{W\ar@<.4ex>[r]&V\ar@<.4ex>[l]},k)$ be a contraction.
Taking the tensor power as in Example~\ref{exa.tensorcoalgebracontractions}, we get a coalgebra contraction
$(\xymatrix{\bar{T}(W)\ar@<.4ex>[r]^\imath&\bar{T}(V)\ar@<.4ex>[l]^\pi},h)$ where
$h=Tk$.
Setting
\[ D_{\partial}=\sum_{n\ge 0}\pi\partial (h\partial)^n\imath=\sum_{n\ge 0}\pi(\partial
h)^n\partial\imath\colon \bar{S}(W)\to \bar{S}(W),\]
then $d+D_{\partial}$ is a codifferential in $\bar{S}(W)$ and
\[ \imath_{\partial}=\sum_{n\ge 0}(h\partial)^n\imath\colon (\bar{S}(W),d+D_{\partial})\to
(\bar{S}(V),d+\partial)\]%
is a morphisms of differential graded coalgebras.
\end{theorem}
\begin{proof}
Since $h(\bigotimes^n V)\subset \bigotimes^n V$ and $\partial(\bigotimes^n V)\subset \bigoplus_{i=1}^{n-1}\bigotimes^i V$ we have
\[ \bigoplus_{i=1}^{n}\bigotimes^i V\subset \ker(\partial h)^n\cap \ker(h\partial)^n\]
and therefore $\partial\in \mathcal{N}(\bar{T}(V),h)$.
According to Proposition~\ref{prop.huebKad} the maps
\[ \imath_{\partial}\colon \bar{T}(W)\to \bar{T}(V),\qquad
D_{\partial}\colon \bar{T}(W)\to \bar{T}(W)\]
are respectively a morphism of graded coalgebras and a coderivation and then
\[ \imath_{\partial}(\bar{S}(W))\subset\bar{S}(V),\qquad
D_{\partial}(\bar{S}(W))\subset\bar{S}(W).\]
The conclusion now follows from Theorem~\ref{thm.relapert}, where
$N=\bar{T}(V)$, $M=\bar{T}(W)$ and $A=\bar{S}(V)$.
\end{proof}
\begin{remark} According to
Proposition~\ref{prop.compositioncompatibility} the construction of Theorem~\ref{thm.symmcoalgebrapert} commutes with composition of contractions.
\end{remark}
\begin{remark} In the notation of Theorem~\ref{thm.symmcoalgebrapert}, if
\[ S^nk\colon \symmetric^n V\to \symmetric^n V,\qquad S^nk=\frac{1}{n!}\sum_{\sigma\in\Sigma_n}\sigma_{{\pmb{\mathtt{tw}}}}\circ T^nk\circ
\sigma_{{\pmb{\mathtt{tw}}}}^{-1},\]
is the symmetrization of $T^nk$ and $Sk=\sum S^nk$, then
$(\xymatrix{\bar{S}(W)\ar@<.4ex>[r]^\imath&\bar{S}(V)\ar@<.4ex>[l]^\pi},Sk)$
is a contraction but in general it is not a coalgebra contraction.
\end{remark}
In the set-up of Theorem~\ref{thm.symmcoalgebrapert} the map $\pi_{\partial}\colon
\bar{T}(V)\to \bar{T}(W)$ is a morphism of graded coalgebras and then induces a morphism
of graded coalgebras $\pi_{\partial}\colon
\bar{S}(V)\to \bar{S}(W)$ such that $\pi_{\partial}\imath_{\partial}$ is the identity on
$\bar{S}(W)$. Unfortunately our proof does not imply that $\pi_{\partial}$ is a morphism of
complexes (unless $(d+\partial)^2=0$ in $\bar{T}(V)$ or $D_{\partial}=0$). However it follows from the homotopy classification of $L_{\infty}$-algebras \cite{K} that a morphism of differential graded coalgebras $\Pi\colon
\bar{S}(V)\to \bar{S}(W)$ such that $\Pi\imath_{\partial}=Id$ always exists.\par
We have proved that the map $\imath_{\partial}\colon \bar{T}(W)\to \bar{T}(V)$ satisfies
the equation $\imath_{\partial}=\imath+(h\partial)\imath_{\partial}$ and then
$\imath_{\partial}\colon \bar{S}(W)\to \bar{S}(V)$ is the unique morphism of symmetric graded coalgebras satisfying the recursive formula
\begin{equation}\label{equ.recursive} \qquad\qquad
p\imath_{\partial}=p\imath+kp\partial\imath_{\partial}\qquad\qquad (\text{where }p\colon \bar{S}(V)\to V \text{ is the projection}).
\end{equation}
It is possible to prove that the validity of the Equation~\ref{equ.recursive} gives a combinatorial description of $\imath_{\partial}$ as sum over rooted trees \cite{cone,fuka} and assures that
$\imath_{\partial}\colon (\bar{S}(W),d+\pi\partial\imath_{\partial})\to (\bar{S}(V),d+\partial)$ is a morphism of differential graded coalgebras (see e.g. the arXiv version of \cite{cone}).
|
\section{Introduction}
\label{sec:introduction}
As computationally demanding research areas expand, the need for
storage space for results and intermediate data increases. Currently
running experiments in areas like high energy physics, atmospheric
science and molecular dynamics already generate petabytes of data
every year. The increasing number of international and even global
scientific collaborations also contributes to the growing need to
share (and protect) data efficiently and effortlessly.
While different research groups have different requirements for a
storage system, a set of key characteristics can be identified. The
storage system needs to be {\it reliable} to ensure data integrity. It
needs to be {\it scalable} and capable to dynamically {\it expand} to
future needs, and given that many research groups today use
computational grids to process their data, it is highly favorable if
the storage system is {\it grid-enabled}.
Lately, the concept of storage clouds has gone from being completely
unknown to being the subject of significant attention from end-users
and developers alike~\cite{scientificcloud}. A storage cloud addresses
many of the needs mentioned above, but from a user perspective it also
hides the complexity of the machinery involved in making the system
work. Such a framework requires well-defined roles for the involved
parties (the Service Providers and Infrastructure Providers). It also
requires these groups to have explicit understandings of and
commitments to their roles. Apart from the conceptual agreements,
there is a fundamental need for sustainable technology on which the
framework can built.
In this paper we will present the design and performance of the
Chelonia storage cloud~\cite{Cheloniasite, CheloniaCloud}. With
Chelonia we aim at designing a system which fulfills the requirements
of global research collaborations, but which also meets the
specifications of a storage cloud, e.g., user-centric interfaces and
transparency. With Chelonia it is possible to build anything from
simple storage systems for sharing holiday pictures to large-scale
storage systems for storing petabytes of scientific data.
This paper is organized as follows: First we give a brief introduction
to distributed data storage solutions in Section~\ref{sec:solutions},
before we in Section~\ref{sec:systemoverview} give an architectural
overview of Chelonia and its services. Section~\ref{sec:operation}
exemplifies important features of Chelonia, while
Sections~\ref{sec:systemperformance} and \ref{sec:systemstability}
present performance and stability of Chelonia, respectively. A
comparison with other storage solutions on the market is given in
Section~\ref{sec:relatedwork}. Section~\ref{sec:futurework} presents
ongoing development work, before the conclusions are given in
Section~\ref{sec:conclusion}.
\section{Distributed Data Storage Solutions}
\label{sec:solutions}
Over the years, different concepts have evolved to deal with the
challenge of handling increasing volumes of data and the fact that the
data tend to be generated and accessed over vast geographic
regions. The largest storage solutions nowadays can be roughly divided
between the data grids developed and used by scientific communities,
and cloud storage arising from the needs of corporate business
communities. While both concepts have the same goal of distributing
large amounts of data across distributed storage facilities, their
focuses are slightly different.
Data grid solutions focus on sharing data stored at several large
storage facilities which are usually supported by public funding and
run by different organizations. A grid storage system provides its
clients with access to data stored at remote storage systems. A
traditional architecture (see e.g.~\cite{datagrid,egeeddm}) typically
consists of an {\it indexing service}, indexing files from storage
resources, a {\it file transfer service} for transferring files, a
{\it replication service} for managing replica locations, and a (often
centralized) {\it metadata catalog} imposing a global namespace on top
of the resources. While they enable the sharing of resources between
large number of users, data grids are often considered to be rather
cumbersome in use and maintenance. For example, there is no common
method of establishing a global namespace and this is achieved only
additional effort of the organization that cares for its own data.
Cloud storage focuses more on providing large amounts of storage to
other organizations, and one cloud storage facility is usually run by
a single organization. The main building blocks of a cloud are the
services. The cloud {\it actors} access the services both to add
resources and utilize resources. The services provides
Quality of Service (QoS) guarantees through Service Level Agreements
(SLA's). In clouds, storage is provided through the concept of Data as
a Service (DaaS), which together with Infrastructure as a Service
(IaaS), Hardware as a Service (HaaS) and Software as a Service (SaaS)
can form a Platform as a Service (PaaS). Hence, by combining services,
the service user can create a customized virtual platform. While this
provides more flexibility for the cloud user and service providers
than the grid concept, it limits the ability of sharing
resources. When a user has set up a virtual computing platform, this
platform is typically limited to be used by this user. Data security
is usually realized through isolating the virtual platform to be used
only by the one user.
Storage clouds are an emerging paradigm, and even though the focus
differs from the data grids, the paradigm has learned from the
experience of the grid paradigm and improved on several features like
usability and payment plans. However, arising from the needs of
corporate business the storage clouds lack features like file-sharing
and high-level tools needed by the scientific communities. The
Chelonia storage cloud is designed to combine the best of two
paradigms for a truly distributed, self-healing, flexible storage
solution, with a replicated metadata catalog, easy-to-use storage
resources and an operating system-agnostic implementation.
\section{Architecture of Chelonia}
\label{sec:systemoverview}
Chelonia consists of a set of SOAP-based web services residing within
the Hosting Environment Daemon (HED)~\cite{HEDdesigndoc} service
container from the ARC grid middleware. Together, the services provide
a self-healing, reliable, robust, scalable, resilient and consistent
data storage system. Data is managed in a hierarchical global
namespace with files and subcollections grouped into
collections\footnote{A concept very similar to files and directories
in most common file systems.}. A dedicated root collection serves as
a reference point for accessing the namespace. The hierarchy can then
be referenced using Logical Names. The global namespace is accessed
the same manner as in local filesystems.
\begin{figure}[htb]
\begin{center}
\includegraphics[width=0.50\columnwidth]{chelonia.eps}
\end{center}
\caption{Schematic of Chelonia's architecture. The figure shows the main
services of Chelonia; The Bartender (cup), the Librarian (book), the
A-Hash (space ship) and the Shepherd (staff). The communication
channels are depicted by black lines.}
\label{fig:fig1}
\end{figure}
Being based on a service-oriented architecture, the Chelonia storage
cloud consists of a set of services as shown in
Figure~\ref{fig:fig1}. The Bartender (cup) provides the high-level
interface to the user; the Librarian (book) handles the entire storage
namespace, using the A-Hash (space ship) as a meta-database; and the
Shepherd (staff) is the front-end for the physical storage node. Note
that any of the services can be deployed in multiple instances for
high availability and load balancing. Before going into the technical
details of Chelonia itself, it may be beneficial to have a brief look
at the middleware providing the communication layer of Chelonia.
\subsection{The Advanced Resource Connector}
\label{sec:arc}
The Chelonia communication layer is provided by the next generation
Advanced Resource Connector (ARC) Grid middleware, developed by
NorduGrid~\cite{NorduGridsite} and the EU-supported KnowARC
project~\cite{KnowARCsite}. The next generation ARC is based upon
web-services which make frequent use of pluggable components for
offering certain capabilities. The ARC services, including Chelonia
(Section~\ref{sec:coreservices}), run inside a container called the
Hosting Environment Daemon (HED).
There are three basic kinds of plugable components for HED: Data
Management Components (DMC's) are used to transfer the data using
various protocols; Message Chain Components (MCC's) are responsible
for the communication within services as well as between
the clients and the services; and Policy Decision Components (PDC's)
are responsible for the security model within the system. In Chelonia,
the Shepherd uses DMC's to transfer files, all client-service and
inter-service communication goes through the SOAP MCC and the
Bartender uses the PDC to decide if a user has access or not.
\subsection{Core Services}
\label{sec:coreservices}
The main components of Chelonia are the four services, the A-Hash, the
Librarian, the Bartender and the Shepherd. They each have separate
roles based on the distinct characteristics of a distributed storage
system. When compared with traditional data grid solutions, the
Librarian may be viewed as an indexing service and the Shepherd as the
manager of the file transfer service. The replication service is
provided by the Shepherd, Librarian and Bartender services acting
together and the Librarian and A-Hash function as a metadata
catalog. When compared with cloud storage solutions, Chelonia provides
DaaS with the Bartender providing a well-defined API for easy-to-use
access. Acting together, the services provide an easy-to-use,
lightweight storage system without single points of failure.
\subsubsection{The A-Hash}
\label{sec:theahash}
The A-Hash service is a database that stores objects which contain
key-value pairs. In Chelonia, it is used to store the global
namespace, all file metadata and information about itself and storage
elements. Being such a central part of the storage system, the A-Hash
needs to be consistent and fault-tolerant. The A-Hash is replicated
using the Oracle Berkeley DB~\cite{berkeley} (BDB), an open source
database library wih a replication API\@. The replication is based on
a single master, multiple clients framework where all clients can read
from the database and only the master can write to the database. While
a single master ensures that the database is consistent at all times,
it raises the problem of having the master as a single point of
failure. If the master is unavailable, the database cannot be updated,
files and entries cannot be added to Chelonia and file replication
will stop working. The possibility of the master failing cannot be
completely avoided, so to ensure high availability means must be taken
to find a new master if the first master becomes unavailable. BDB uses
a variant of the Paxos algorithm~\cite{Paxos} to elect a master
amongst peer clients: Every database update is assigned an increasing
number. In the event of a master going offline, the clients sends a
request for election, and a new master is elected amongst the clients
with the highest numbered database update.
\subsubsection{The Librarian}
\label{sec:librarian}
The Librarian service manages the hierarchy and metadata of files and
collections, handles the Logical Names and monitors the Shepherd
services. The Librarian service is stateless, instead it stores all
the persistent information in the A-Hash. This makes it possible to
deploy any number of independent Librarian services to provide
high-availability and load-balancing. In this case all the Librarians
should communicate with the same set of A-Hashes in order to use the
same database of metadata. As only the master A-Hash can be written to
and the Librarian cannot know {\it a priori} which A-Hash replica is
the master, the Librarian needs to get this information from one of
the A-Hashes. For this reason, the master A-Hash stores the list of
all available A-Hashes, so that the information is
replicated to all A-Hash replicas. As all A-Hash replicas are
readable, the Librarian only needs to know about a few of the A-Hashes
at start-up to be able to get this list. During run-time the Librarian
holds a local copy of the A-Hash list and refreshes it both regularly
and in the case of a failing connection.
\subsubsection{The Shepherd}
\label{sec:shepherd}
Each instance of the Shepherd service manages a particular storage
node and provides a uniform interface for storing and accessing file
replicas. On a storage node there must be at least one independent
storage element service (with an interface such as HTTP(S), ByteIO,
etc.)\ which performs the actual file transfer. A storage node then
consists of a Shepherd service and a storage element service connected
together. Storage element services can either be provided by ARC or by
third-party services. For each kind of storage element service, a
Shepherd backend module is needed to enable the Shepherd service to
communicate with the storage element service, e.g., to initiate file
uploads, downloads and removal, and to detect whether a file transfer
was successful or not. Currently there are three Shepherd backends:
One for the ARC native HTTP(S) server called Hopi; one for the Apache
web server; and one for a service which implements a subset of the
ByteIO interface. In addition to storing files and providing access to
them, the Shepherd is responsible for checking if a file replica is
valid and, if necessary, initiating replication of the file to other
Shepherds.
\subsubsection{The Bartender}
\label{sec:bartender}
The Bartender service provides a high-level interface of the storage
system for the clients (other services or users). The clients can
create and remove collections (directories), create, get and remove
files, and move files and collections within the namespace using
Logical Names. Access policies associated with files and collections
are evaluated by the Bartender (using the PDC plugin of HED) every
time a user wants to access them. The Bartender communicates with the
Librarian and Shepherd services to execute the client's requests. The
file content itself does not go through the Bartender; file transfers
are directly performed between the storage nodes and the clients.
The Bartender also supports so-called gateway modules which make it
possible to communicate with third-party storage solutions, thus
enabling the user to access multiple storage systems through a single
Bartender client. These modules are protocol-oriented in the sense
that external storage managers which support a certain protocol will
be handled using the gateway module based on that protocol. While
excluding some of the features provided by accessing storage managers
directly, this approach reduces the number of gateway modules
required for different storage managers. The currently available
gateway module is based on the GridFTP protocol\cite{gridftp}.
\subsection{Security}
\label{sec:security}
As is the case for all openly accessible web services, the security
model is of crucial importance for the Chelonia storage cloud. While
the security of the communication with and within the storage system
is realized through HTTPS with standard X.509 authentication, the
authorization related security architecture of the storage can be
split into three parts; the inter-service authorization; the
transfer-level authorization; and the high-level authorization:
\begin{itemize}
\item The inter-service authorization maintains the integrity of the
internal communication between services. There are several
communication paths between the services in the storage system. The
Bartenders send requests to the Librarians and the Shepherds, the
Shepherds communicate with the Librarians and the Librarians with
the A-Hash. If any of these services are compromised or a new rogue
service gets inserted in the system, the security of the entire
system is compromised. To enable trust between the services, they
need to know each other's Distinguished Names (DN's). This way a
rogue service would need to obtain a certificate with that exact DN
from some trusted Certificate Authority (CA).
\item The transfer-level authorization handles the authorization in
the cases of uploading and downloading files. When a transfer is
requested, the Shepherd will provide a one-time Transfer URL (TURL)
to which the client can connect. In the current architecture, this
TURL is world-accessible. This may not seem very secure at
first. However, provided that the TURL has a very long, unguessable
name, that it is transfered to the user in a secure way and that it
can only be accessed once before it is deleted, the chance of being
compromised is fairly low.
\item The high-level authorization considers the access policies for
the files and collections in the system. These policies are stored
in the A-Hash, in the metadata of the corresponding file or collection,
providing a fine-grained security in the system.
\end{itemize}
\subsection{Accessing Chelonia}
\label{sec:clienttools}
Being the only part a user will (and should) see from a storage
system, the client tools are an important part of the Chelonia storage
cloud. In addition to a vanilla command-line interface, two ways
of accessing Chelonia are supported.
\subsubsection{FUSE Module}
The FUSE module provides a high-level access to the storage
system. Filesystem in Userspace (FUSE)~\cite{FUSE} provides a simple
library and a kernel-userspace interface. Using FUSE and the ARC
Python interface, the FUSE module allows users to mount the storage
namespace into the local namespace, enabling the use of graphical file
browsers as shown in the screenshot in
Figure~\ref{fig:fusescreenshot}.
\begin{figure}[htb]
\begin{center}
\includegraphics[width=0.70\columnwidth]{FUSE.eps}
\end{center}
\caption{Screenshot of the Chelonia FUSE module in use. Through the
FUSE module Chelonia offers users a drag and drop functionality to
upload or download files to the storage cloud.}
\label{fig:fusescreenshot}
\end{figure}
\subsubsection{Grid Job Access}
To access data through the ARC middleware client tools, one needs to
go through Data Manager Components (DMC's). These are
protocol-specific plugins to the client tools. For example, to access
data from a HTTPS service, the HTTP DMC will be used with a URL
starting with \verb!https://!, to access data from an SRM service, the
SRM DMC will be used with a URL starting with
\verb!srm://!. Similarly, to access Chelonia, the ARC DMC will be used
with a URL starting with \verb!arc://!.
The ARC DMC allows grid jobs to access Chelonia directly. As long as
A-REX, the job execution service of ARC, and ARC DMC are installed on
a site, files can be both downloaded and uploaded by specifying the
corresponding URL's in the job description. In this case, the
Bartender URL needs to be embedded in the URL as a URL option. For
example, if a job requires an input file \verb!/user/me/input.dat!,
the URL specified in the job description will be as follows (given
that the file can be found by a Bartender with URL
\verb!https://storage/Bartender!):
\begin{verbatim}
arc:///user/me/input.dat?BartenderURL=https://storage/Bartender
\end{verbatim}
\section{Chelonia in Operation}
\label{sec:operation}
Both the Chelonia storage system and its clients can be installed from
binary packages (available for several different platforms) or after
compiling the source packages. A fully operational storage cloud
requires a minimum installation of one instance of every service
described above. The Chelonia Administrator
manual~\cite{cheloniaadmin} gives detailed instructions on how
to install, configure and run the services. In order for users to
interact with Chelonia, several user tools are provided. These are
documented both in the Chelonia user
manual~\cite{cheloniausermanual} and Linux man pages and directly through
command line calls.
For the user, transfering files to and from Chelonia are simple
operations. For example, if a user wants to upload a file
\verb!orange.jpg! to Chelonia, he/she can use, e.g., the Chelonia
CLI. Assuming that the URL of one or more Bartenders and the required
number of file replicas are written in a configuration file, the user
gives the command
\begin{verbatim}
chelonia put orange.jpg /user/me/orange.jpg
\end{verbatim}
Note that there is
no need for the user to know where files are physically stored or will
be stored in Chelonia.
Under the hood of Chelonia, the Bartender receiving the request from
the CLI contacts a Librarian to create an entry in the Chelonia
namespace. If the Librarian confirms the new entry, the Bartender then
contacts a Shepherd to get a transfer URL which is returned to the
CLI. When the CLI has uploaded the file, the Shepherd queries the
Librarian to find out how many replicas are needed and, if needed,
initiates a file transfer to another Shepherd. In the case of
downloading a file, the Bartender gets the locations of the file from
a Librarian, chooses one of them and contacts the corresponding
Shepherd for a transfer URL.
When in operation, the Chelonia storage cloud is a pulsing system where
heartbeats are periodically sent from each Shepherd to a Librarian
together with information about replicas whose state changed since the
last heartbeat. Heartbeats are stored in the A-Hash, thus making them
visible to all Librarians in the system. If any of the Librarians
notices that a Shepherd is late with its heartbeat, it will mark all
the replicas in that Shepherd as offline.
In addition to the heartbeat, the Shepherds periodically check with
the Librarians to see if there are sufficient replicas of the files in
Chelonia and if the checksums of the replicas are correct. If a file
is found to have too few replicas, the Shepherd informs a Bartender
about this situation and together they ensure that a new replica is
created at a different storage node. A file having too many replicas
will also be automatically corrected by Chelonia - the first Shepherd
to notice this will mark its replica(s) as unneeded and later delete
it (them). Replicas with invalid checksums are marked as invalid, and
as soon as possible replaced with a valid replica.
With the Chelonia gateway module, a user can mount external storage
systems into the Chelonia namespace. For example, if a Chelonia user has
access to a set of files stored in dCache~\cite{dCache} (see
Section~\ref{sec:relatedwork}) under \verb!/fruits! he can add a mount
point (say \verb!/my/dCache!) to easily access these data
through the Chelonia namespace, i.e., using standard Chelonia commands
like
\begin{verbatim}
chelonia get /my/dCache/fruits/apple.jpg
\end{verbatim}
More technically, when the Bartender looks up the entry
\verb!/my/dCache! which is a mount point to dCache, it will use the
corresponding gateway module to generate an external URL which the
client tool will use to contact dCache directly. This way, Chelonia
can include third-party storage namespaces in its global namespace by
simply storing a single entry.
\section{System Performance}
\label{sec:systemperformance}
\subsection{Adding and Querying the Status of Files}
\label{sec:addingfiles}
In a hierarchical file system files are stored in levels of
collections and sub-collections. The time to add or get a file depends
mainly on two factors; the number of entries in the collection where
the file is inserted, and the number of parent collections to the
collection where the file is inserted (the depth of the
collection). Based on these two factors we have run two different
tests:
\begin{itemize}
\item {\bf Depth test} tests the performance when creating many levels
of sub-collections. The test adds a number of sub-collections to a
collection, measures the time to add and stat the sub-collections,
then adds a number of sub-sub-collections to one of the
sub-collections and so forth. To query a collection at a given level
means that all the collections at the lower levels needs to be
queried first. In Chelonia, each query causes a message to be sent
through TCP\@. Hence, it is expected that time will increase
linearly as the level of collections increases. As every message is
of equal size, this test ideally depends only on network latency.
\item {\bf Width test} tests the performance when adding many entries
(collections) to one collection. The test is carried out by adding a
given number of entries to a collection and measuring the time to add
each entry and the time to stat the created entry. When adding an
entry to a collection, the system needs to check first if the entry
exists. In Chelonia, this means that the list of entries in the
collection needs to be transferred through TCP\@. It is therefore
expected that the time to add an entry will increase linearly as
this list increases and ideally the time will depend only on the
bandwidth of the network.
\end{itemize}
Both tests were run in two types of environments: In the Local Area
Network (LAN) setup four computers were connected to the same switch.
A centralized A-Hash service, a Librarian, a Bartender, and the client
were each run on a separate computer. In the Wide Area Network (WAN)
setup, the client and Bartender were located in Uppsala, Sweden, while
the Librarian and a replicated A-Hash consisting of three replicas were
located in Oslo, Norway.
\begin{figure}[ht]
\centering
\subfigure[Services running on LAN]{
\includegraphics[width=0.47\columnwidth]{DepthTest_LAN.eps}
\label{fig:depthTestLan}
}
\subfigure[Services running on WAN]{
\includegraphics[width=0.47\columnwidth]{DepthTest_WAN.eps}
\label{fig:depthTestWan}
}
\caption{Time to add an entry to a collection (continuous line) and
time to get status of a collection (dashed line) given the
hierarchical depth of the collection. }
\end{figure}
The test result for the depth test is shown in Figures
\ref{fig:depthTestLan} (LAN) and \ref{fig:depthTestWan} (WAN)\@. The
continuous lines show the time to create an entry, while the dashed
lines show the time to get the status of the collection. All the plots
are averages of five samples, with the error bars representing the
minimum and maximum values. The LAN test shows a near-perfect linear
behaviour, with the error bars too small to be seen. As mentioned
earlier, since the packet size for each message is constant in this
test, the main bottleneck (apart from Chelonia itself) is the network
latency. Since all computers in the LAN test are connected to the same
switch we can assume that the latency is near constant. Hence, the LAN
test shows that in a very simple network scenario Chelonia works as
expected, with the network being the major bottleneck. Notice also
that creating an entry consistently takes 0.021~s longer than
getting the status of the collection, again corresponding to the extra
message needed to create an entry.
In the WAN test, the complexity is a bit increased, as in addition to
sending messages over WAN, the A-Hash is now replicated. The time
still increases linearily, albeit with more fluctuation due to the WAN
environment. Creating an entry at the first level in the hierarchy now
takes 0.11~s longer than getting the status of the collection,
corresponding to the fact that the entry needs to be replicated three
times. However, at higher levels, getting the status over WAN is
actually faster than getting the status over LAN. The effect of the
replicated A-Hash will be discussed in more detail in Section
\ref{sec:ahashtest}.
\begin{figure}[ht]
\centering
\subfigure[Services running on LAN]{
\includegraphics[width=0.47\columnwidth]{WidthTest_LAN.eps}
\label{fig:widthTestLan}
}
\subfigure[Services running on WAN]{
\includegraphics[width=0.47\columnwidth]{WidthTest_WAN.eps}
\label{fig:widthTestWan}
}
\caption{Time to add an entry to a collection (continuous line) and
time to get status of a collection (dashed line) given the
number of entries already in the collection.}
\end{figure}
The test results for the width test are shown in Figures
\ref{fig:widthTestLan} (LAN) and \ref{fig:widthTestWan} (WAN)\@. The
continous lines show the time to add entries to a collection and the
dotted lines show the time to get the status of a collection
containing the given number of entries. The operations were repeated
five times, with the plots showing the averages of these five
samples. As expected, the time increase linearily with increasing
number of entries. The results of the WAN test fluctuate more than
those of the LAN test, which is to be expected; in the LAN test all
services are on computers connected to the same switch, while in the
WAN test, the services are distributed between two different
countries. However, the WAN test shows similar linearity, albeit with
a slightly higher response time. It is worth noticing that for the
width test, in contrast to the depth test, we see no benefit of using
a replicated A-Hash. This is due to the fact that the bandwith is the
limiting factor in this test.
An interesting feature of both tests is that while creating an entry
takes more time when there are only few entries in the collection,
creating new entries is actually faster than stating the collection
when the collection has many entries. This is due to the fact that
getting the status of a collection requires the metadata of the
collection (and hence the list of entries in the collection) to be
transfered first from the A-Hash to the Librarian, and second from the
Librarian to the Bartender and last from the Bartender to the
client. When creating an entry, neither the Bartender nor the client
needs this list of entries, so that less data is transfered between
services. However, for fewer entries, creating new entries is more
expensive since the Bartender needs to query the Librarian twice,
first to check if it is allowed to add the entry and second to
actually add it.
\subsection{File Replication}
\label{sec:filereplication}
The concept of automatic file replication in Chelonia was presented in
Section~\ref{sec:operation}. In this section we will demonstrate both
how Chelonia works with different file states to ensure that a file
always has the requested number of valid (ok) replicas and how
Chelonia distributes the replicas in the system in order to ensure
maximum fault tolerance of the system.
The test system consists of one Bartender, one Librarian, two A-Hashes
(one client and one master) and five Shepherds. All services are
deployed within the same LAN. As a a starting point 10 files of 114~MB
are uploaded to the system and for each file 4 replicas are
requested. Thus the initial setup of the test system contains 40
replicas with an initial distribution of file replicas as shown in
Table~\ref{tab:initialfiledistribution}.
\begin{table}[ht]
\centering
\begin{tabular}{c r r }
\hline
Shepherd & Initial & Final\\
\hline
\hline
S1 & 9 & 9\\
S2 & 8 & 7\\
S3 & 8 & 8\\
S4 & 7 & 9\\
S5 & 8 & 7\\
\hline
Total & 40 & 40\\
\end{tabular}
\caption{Initial and final load distribution of 40 files on 5 Shepherds}
\label{tab:initialfiledistribution}
\end{table}
The first phase of the file replication test was to kill one of the
Shepherd services, S3, of the test system. Chelonia soon recognized
the loss of this service (no heartbeat received within one cycle) and
started compensating for the lost replicas. File replicas in Chelonia
have states ALIVE, OFFLINE, THIRDWHEEL or CREATING which are recorded
in the A-Hash. Initially the test system had 40 ALIVE replicas (10
files with 4 replicas each), but when the Librarian did not get the S3
heartbeat it changed the state of the 8 replicas stored in S3 to
OFFLINE.
At this point a number of files stored in our Chelonia setup had too
few ALIVE replicas. As explained above, the Shepherds check
periodically that files with replicas stored on its storage element
have the correct number of ALIVE replicas. Thus, in the next cycle the
S1, S2, S4 and S5 Shepherds started to create new file replicas which
initially appeared in the system with the state CREATING.
Figure~\ref{fig:FileReplication} gives a graphical overview of the
number of replicas and the replica states in the test system every 15
seconds. At 15~s (00:15) all 40 file replicas were ALIVE as explained
above, but soon thereafter S3 was turned off and the other Shepherds
started creating new replicas. Queried at 30~s (00:30) the system
contained 32 ALIVE, 8 OFFLINE and 1 CREATING file replica.
While the system worked on compensating for the loss of S3, the second
phase of the replication test was initiated by turning the S3 shepherd
online again. The reappearence of the S3 Shepherd can be seen in
Figure~\ref{fig:FileReplication} at 90s (01:30) as a significant
increase in the number of ALIVE replicas. In fact there were now too
many ALIVE replicas in the system and at 105~s (01:45) 2 replicas were
marked as THIRDWHEEL (the Chelonia state for redundant
replicas). THIRDWHEEL replicas are removed from Chelonia as soon as
possible, and during the next 45~s the system removed all such
replicas. A query of the system at 165~s (02:45) shows that the system
once again contained 40 ALIVE replicas. The final distribution of
replicas between Shepherds is given in
Table~\ref{tab:initialfiledistribution}.
\begin{figure}[htb]
\begin{center}
\includegraphics[width=0.8\columnwidth]{ReplicationTest.eps}
\end{center}
\caption{Number of replicas and their corresponding states when
querying the test system every 15~seconds. Shepherd S3 is turned
offline between 00:15 and 00:30 and back online between 01:15 and
01:30.}
\label{fig:FileReplication}
\end{figure}
\subsection{Multi-User Performance}
\label{sec:multiuser}
While any distributed storage solution must be robust in terms of
multi-user performance, it is particularly important for a
grid-enabled storage cloud like Chelonia where hundreds of grid jobs
and interactive users are likely to interact with the storage system
in parallel. To analyze the performance of Chelonia in such
environments we have studied the response time of the system while
increasing the number of simultaneous users (multi-client).
\begin{figure}[ht]
\centering
\includegraphics[width=0.8\columnwidth]{MultiClient_LAN.eps}
\label{fig:MultiClientLan}
\caption{Average (square), minimum (lower bar) and maximum (bar)
system response time as a function of the number of simultaneous
clients of the system. Each client creates 50 collections
sequentially.}
\end{figure}
Due to limited hardware resources, multiple clients for the tests were
simulated by running multiple threads from three different
computers. Each client thread creates 50 collections sequentially and
tests were done for an increasing number of simultaneous clients. For each
test the minimum, average and maximum time used by the client was
recorded. Figure~\ref{fig:MultiClientLan} shows the system response
times for up to 100 simultaneous clients using the above-mentioned
testbed deployment. The shown test was run in a LAN environment with
one centralized A-Hash, one Librarian and one Bartender.
The test results show that the response time of the system increases
linearly with an increasing number of simultaneous clients. From 40
clients and onwards the difference between the fastest client and
slowest client starts to become sizeable. When running 50 clients or
more in parallel, it was occasionally observed that a client's request
failed due to the heavy load of the system. When this happened, the
request was retried until successfully completed, as shown by the
slightly fluctuating slope of the mean curve. The same linearity was
seen in the corresponding WAN test (not shown), albeit with a factor
two higher average time, consistent with the results observed in the
depth and width tests in Section~\ref{sec:addingfiles}.
As can be noted in Figure~\ref{fig:MultiClientLan}, for more than 30
clients, the maximum times increase approximately linearly while the
minimum times are close to constant. The reason for this is a
limitation on the number of concurrent threads in the Hosting
Environment Daemon (HED). If the number of concurrent requests to HED
reaches the threshold limit, the requests are queued so that only a given
number of requests are processed at the same time. In the test each
client used only one connection for creating all 50 collections. Hence,
the fastest request was one that had not been queued so that when
the number of requests was above the threshold, the minimum timing did
not depend on the total number of clients. On the other hand, the
slowest request was queued behind several other requests so
that the maximum time increased with the number of simultaneous
clients.
\subsection{Centralized and Replicated A-Hash}
\label{sec:ahashtest}
As the A-Hash stores all metadata about files, file locations and
shepherds, it is important that the A-Hash is fault tolerant and able
to survive even fatal hardware failures. While in theory replicating
the A-Hash provides these features, the replication adds complexity to
the A-Hash in that all data need to be replicated to all A-Hash
instances. Additionally, in the event of a failing A-Hash instance, the
Librarians need to seamlessly find and connect to other A-Hashes.
To test the fault tolerance and performance overhead of the replicated
A-Hash in a controlled environment, four tests have been set up
with services and a client on different computers in the same LAN:
\begin{enumerate}
\item {\bf Centralized}: One client contacting a centralized A-Hash
was set up as a benchmark, as this is the simplest possible
scenario.
\item {\bf Replicated, stable}: One client contacting three A-Hash
instances (one A-Hash master, two A-Hash clients) randomly. All the
A-Hashes were running during the entire test.
\item {\bf Replicated, unstable clients}: Same setup as in point 2,
but with a random A-Hash client restarted every 60 seconds.
\item {\bf Replicated, unstable master}: Same setup as in point 2,
but with the master A-Hash restarted every 60 seconds.
\end{enumerate}
While setups 1 and 2 test the differences in having a centralized
A-Hash and a replicated A-Hash, setups 3 and 4 tests how the system
responds to an unstable environment. In all four setups the system
has services available for reading at all times. However, in setup 3
one may need to reestablish connection with an A-Hash client and in
setup 4 the system is not available for writing during the election of
a new master. During the test the
client computer constantly and repeatedly contacted the A-Hash for
either writing or reading for 10 minutes. During write tests the client
computer reads the newly written entry to ensure it is correctly
written.
\begin{table}[ht]
\centering\small
\begin{tabular}{l |c | c | c}
\hline
\multicolumn{4}{c}{\bf Reading}\\
\hline
& {\it Minimum (s)} & {\it Average (s)} &
{\it Maximum (s)}\\
\hline
Centralized & 0.003399 & 0.003780 & 0.013441 \\
Replicated, stable & 0.003453 & 0.003738 & 0.013261\\
Replicated, unstable clients & 0.003412 & 0.003754 & 0.289535\\
Replicated, unstable master & 0.003402 & 0.003763 & 1.971131\\
\hline
\multicolumn{4}{c}{}\\\hline
\multicolumn{4}{c}{\bf Writing}\\
\hline
& {\it Minimum (s)} & {\it Average (s)} & {\it Maximum (s)}\\
\hline
Centralized & 0.003828 & 0.004260 & 0.014459 \\
Replicated, stable & 0.016866 & 0.033902 & 1.057602 \\
Replicated, unstable clients & 0.016434 & 0.034239 & 1.131142 \\
Replicated, unstable master & 0.016293 & 0.044868 & 60.902862\\
\hline
\end{tabular}
\caption{Timings for reading from and writing to a centralized A-Hash
compared with a stable replicated A-Hash, a replicated A-Hash where
clients are restarted and a replicated A-Hash where the master is
restarted.}
\label{tab:ahashreadwright}
\end{table}
Table~\ref{tab:ahashreadwright} shows timings of reading from and
writing to the A-Hash for the four setups. As can be seen, for
reading, all four setups have approximately the same
performance. Somewhat surprisingly, the replicated setups actually
perform better than the centralized setup, even though only one client
computer was used for reading. While reading from more A-Hash
instances is an advantage for load balancing in a multi-client
scenario, one client computer can only read from one A-Hash instance
at a time. Thus, the only difference between the centralized and
replicated setups in terms of reading is how entries are looked up
internally in each A-Hash instance, where the centralized A-Hash uses
a simple Python dictionary, while the replicated A-Hash uses the
Berkeley DB which has more advanced handling of cache and memory.
Looking at the write performance in Table~\ref{tab:ahashreadwright},
there is a more notable difference between the centralized and
the replicated setups. On average, writing to the replicated A-Hash
takes almost 10 times as long as writing to the centralized
A-Hash. The reason for this is that the master A-Hash will not
acknowledge that the entry is written until all the A-Hash instances
have confirmed that they have written the entry. While this is rather
time-consuming, it is more important that the A-Hash is consistent
than fast. It is however worth noticing when implementing services
using the A-Hash, that reading from a replicated A-Hash is much faster
than writing to it.
\section{System Stability}
\label{sec:systemstability}
While optimal system performance may be good for the day-to-day user
experience, the long-term stability of the storage system is an
absolute requirement. It does not help to have a response time of a
few milliseconds under optimal conditions if the services need to be
frequently restarted due to memory leaks or if the servers become
unresponsive due to heavy load.
To test the system stability, a Chelonia deployment was run
continuously over a week's time. During the test a client regularly
interacted with the system, uploading and deleting files and listing
collections. The deployment consisted of one Bartender, one Librarian,
three A-Hashes and two storage nodes, each consisting of a Shepherd
and a Hopi service. All the services and the client ran on separate
servers in a LAN environment.
Figure~\ref{fig:RSSconsumption} shows the overall memory utilization
for seven of the services for the entire run time (top), the A-Hashes and
the Librarian in the first 24 hours (bottom left) and the Librarian,
the Bartender and one of the Shepherds in the last 37 hours (bottom
right). The memory usage was measured by reading the memory usage of
each service process in 5 seconds intervals using the Linux \verb!ps!
command.
\begin{figure}
\includegraphics[angle=90,width=0.9\columnwidth]{RSSconsumption.eps}
\caption{Resident memory utilization of the Chelonia services during
an 8 day run.}
\label{fig:RSSconsumption}
\end{figure}
The most crucial part of Chelonia, when it comes to handling server
failures, is the replicated A-Hash. If the A-Hash becomes unavailable,
the entire system is unavailable. If a client A-Hash goes down, the
Librarians may need to find a new client. If the master A-Hash goes
down, the entire system will be unavailable until a new master is
elected. The bottom left of Figure~\ref{fig:RSSconsumption} shows the
memory consumption of the three A-Hashes, A-Hash2, A-Hash3 and
A-Hash7, and the Librarian during the first 24 hours of the stability
test. When the test started A-Hash3 (red line) was elected as A-Hash
master, and the Librarian started using A-Hash7 (blue line) for read
operations. After 2.5 hours, A-Hash3 was stopped (seen by the sudden
drop of the red line), thus forcing the two remaining A-Hashes to
elect a new master between them. While not visible on the figure, the
A-Hash, and hence Chelonia, was unavailable for a 10 seconds period
during the election, which incidentally was won by A-Hash2 (black
line). After three additional hours, A-Hash3 was restarted, thus
causing an increase of memory usage for the master A-Hash as it needed
to update A-Hash3 with the latest changes in the database. Eight hours
into the test, the same restart procedure was carried out on A-Hash7
which was connected to the Librarian. This time there was no
noticeable change in performance. However A-Hash3 increased memory
usage when the Librarian connected to it.
\begin{figure}
\includegraphics[width=1.1\columnwidth]{CPUConsumption.eps}
\caption{ CPU load of the Chelonia services during an 8 day run. Each
point is an average of the CPU usage of the previous 60 seconds.}
\label{fig:CPUconsumption}
\end{figure}
The bottom right of Figure~\ref{fig:RSSconsumption} shows the memory
usage of the Bartender, the Librarian and one of the shepherds in the
last 37 hours of the test. Perhaps most noteworthy is that the memory
usage is very stable. The main reason for this is due to the way
Python, the programming language of Chelonia, allocates memory. As
memory allocation is an expensive procedure, Python tends to allocate
slightly more memory than needed and avoids releasing the already
acquired memory. As a result, the memory utilization gets evened out
after a period of time even though the usage of the system
varies. During the run-time of the test, files of different sizes were
periodically uploaded to and deleted from the system. The slight jump
in memory usage for the Bartender (green line) was during an
extraordinary upload of a set of large files. This jump was followed
by an increase in memory for the Librarian when the files were
starting to be replicated between the Shepherds, thus causing extra
requests to the Librarian.
Figure~\ref{fig:CPUconsumption} shows the CPU load for six of the
services. While the load on the A-Hashes, the Bartender and the
Librarian are all below 2.5\% of the CPU, Shepherd-1 (bottom left) and
Shepherd-8 (bottom right) use around 10\% and 20\%,
respectively. While the difference between the Shepherds is due to the
fact that Shepherd-1 was run on a server with twice the number of
CPU's as the server of Shepherd-8, the difference between the
Shepherds and the other services is due to the usage pattern during
the test. To confirm that the files stored on the storage node are
healthy, the Shepherd calculates a checksum for each file, first when
the file is received and later periodically. In periods where no new
files are uploaded, the Shepherds use almost no CPU as already stored
files don't need frequent checksum calculations. However, when files
are frequently uploaded, deleted and re-uploaded, as was the case
during the test, the number of checksum calculations, and hence the
CPU load, increases significantly. This can particularly be seen on
January 1 and 4 when a set of extra large files were uploaded, causing
spikes in the CPU load of the two Shepherds. Note also that the spikes
occurs at the same time for both Shepherd, as should be expected in a
load balanced system.
\section{Related Work}
\label{sec:relatedwork}
There are a number of grid and cloud storage solutions on the market,
focused on different storage needs. While direct performance
comparison with Chelonia is beyond the scope of this paper, some
similarities and differences between Chelonia and related storage
solutions are worth mentioning.
In the cloud storage family, Amazon Simple Storage Service
(S3)~\cite{amazons3} promises unlimited storage and high
availability. Amazon uses a two-level namespace as opposed to the
hierarchical namespace of Chelonia. In the security model of S3, users
have to implicitly trust S3 entirely, whereas in Chelonia users and
services need to trust a common independent third party Certificate
Authority. Additionally, S3 lacks fine-grained delegation and access
control lists are limited to 100 principals, limiting the usability
for larger scientific communities~\cite{amazons3forsciencegrids}.
While Chelonia is designed for geographically distributed users and
data storage, Hadoop \cite{hadoop} with its file system HDFS is
directed towards physically closely-grouped clusters. HDFS builds on
the master-slave architecture where a single NameNode works as a
master and is responsible for the metadata whereas DataNodes are used to
store the actual data. Though similar to Chelonia's metadata service,
the NameNode cannot be replicated and may become a bottleneck in the
system. Additionally, HDFS uses non-standard protocols for
communication and security while Chelonia uses standard protocols like
HTTP(S), GridFTP and X509.
When compared to typical grid distributed data management solutions,
the closest resemblance with Chelonia is the combination of the
storage element Disk Pool Manager (DPM) and the file catalog
LCG~\footnote{LHC (Large Hadron Collider) Computing Grid} File
Catalog (LFC)~\cite{dpmlfc}. By registering all files uploaded to
different DPM's in LFC one can achieve a single uniform namespace
similar to the namespace of Chelonia. However, where Chelonia has a
strong coupling between the Bartenders, Librarians and Shepherds to
maintain a consistent namespace, DPM and LFC have no coupling such
that registration and replication of files is handled on the client
side. If a file is removed or altered in DPM, this may not be
reflected in LFC. In Chelonia, a change of a file has to be registered
through the Bartender and propagated to the Librarian before it is
uploaded to the Shepherd.
dCache~\cite{dCache} differs from Chelonia in that dCache has a
centralized set of core services while Chelonia is distributed by
design. dCache is a service-oriented storage system which combines
heterogeneous storage elements to collect several hundreds of
terabytes in a single namespace. Originally designed to work on a
local area network, dCache has proven to be useful also in a grid
environment, with the Nordic Data Grid Facility (NDGF) dCache
installation~\cite{DSSWithdCache} as the largest example. There, the
core components, such as the metadata catalogue, indexing service and
protocol doors are run in a centralized manner, while the storage
pools are distributed. Chelonia, designed to have multiple instances
of all services running in a grid environment, will not need a
centralized set of core services. Additionally, dCache is relatively
difficult to deploy and integrate with new applications. Being a more
light-weight and flexible storage solution, Chelonia aims more towards
new, less demanding, user groups which are generally less familiar
with grid solutions.
Scalla \cite{Scalla} differs from Chelonia in that Scalla is designed
for use on centralized clusters, while Chelonia is designed for a
distributed environment. Scalla is a widely used software suite
consisting of an xrootd server for data access and an olbd server for
building scalable xrootd clusters. Originally developed for use with
the physics analysis tool ROOT \cite{root} , xrootd offers data access
both through the specialized xroot protocol and through other
third-party protocols. The combination of the xrootd and olbd
components offers a cluster storage system designed for low latency,
high bandwidth environments. In contrast, Chelonia is optimized for
reliability, consistency and scalability at some cost of latency and
is more suitable for the grid environment where wide area network
latency can be expected to be high.
Unlike Chelonia, iRODS \cite{irods} does not provide any storage
itself but is more an interface to other, third-party storage
systems. Based on the client-server model, iRODS provides a flexible
data grid management system. It allows uniform access to heterogeneous
storage resources over a wide area network. Its functionality, with a
uniform namespace for several Data Grid Managers and file systems, is
quite similar to the functionality offered by our gateway
module. However, iRODS uses a database system for maintaining the
attributes and states of data and operations. This is not needed with
Chelonia's gateway modules.
\section{Future Work}
\label{sec:futurework}
In addition to the continous process of improvements and
code-hardening (based on user feedback) there are plans to add some
new features.
The security of the current one-time URL based file transfers could be
improved by adding to the URL a signed hash of the IP and the DN
of the user. In this way the file transfer service could do
additional authorization, allowing the file transfer only for the same
user with the same IP.
Because of the highly modular architecture of both Chelonia and the
ARC HED hosting environment, the means of communication between the
services could be changed with a small effort. This would enable less
secure but more efficient protocols to replace HTTPS/SOAP when
Chelonia is deployed inside a firewall. This modularity also allows
additional interfaces to Chelonia to be implemented easily. For
example, an implementation of the WebDAV protocol would make the
system accesible to standard clients built into the mainstream
operating systems.
Another possible direction for enhancing the functionality of Chelonia
is to add handling of SQL databases. In addition to files, the system
could store database objects and use databases as storage nodes to
store them. SQL databases allow running extensive queries to get the
desired information. In a distributed environment, high availability
and consistency is often ensured by the replication of data. Access to
multiple copies of the data in the system also allows queries to be
run in parallel. Consistent, multiple copies of the data also provdies
a simple, transparent platform for scalable access to the same data to
a large number of distributed clients.
\section{Conclusions}
\label{sec:conclusion}
Chelonia is a cloud-like storage solution with grid
capabilities. While its core services resembles those of a traditional
data grid, the single-entry interface and the capabilities resemble
those of storage clouds.
An important part of developing a distributed storage system is proper
testing of the system, both in terms of performance and stability. The
presented tests are designed to give an understanding of how Chelonia
behaves in a real-life environment while at the same time controlling
the environment enough to get interpretable results. The tests have
shown that Chelonia can handle both deep and wide hierarchies as
expected, both in LAN environments and in WAN environments. The system
has shown self-healing capabilities, both in terms of individual
service stops and in terms of file availability. Multiple clients have
accessed the system simultaneously with reasonable performance results
and, even more importantly, Chelonia has been heavily used for more
than a week with stable performance even with vital services being
shut down during the test.
\section{Acknowledgements}
\label{Acknowledgements}
We wish to thank Mattias Ellert for vital comments and
proof reading. Additionally, we would like to thank UPPMAX, NIIF and
USIT for providing resources for running the storage tests.
The work has been supported by the European Commission through the KnowARC
project (contract nr. 032691) and by the Nordunet3 programme through the NGIn
project.
|
\section{Introduction}
A fundamental result in the thermodynamics of computation is
Landauer's erasure principle (LEP), which places a lower bound on
the average heat dumped to a bath when the information in a memory
device is erased. The principle holds for processes that satisfy
two essential features: The process is carried out in the same way
independent of the information stored in the device and it
restores the device to a known \emph{standard state} at the end.
These processes are usually known as the \emph{Landauer erasure}
processes. Landauer has shown that if the device holds $I$ bits of
information, then during any Landauer erasure, the average amount
of heat that must be dumped to the bath is at least $I k_B T \ln 2
$ where $k_B$ is Boltzmann's constant and $T$ is the temperature
of the bath \cite{landauer}.
The principle lies at the heart of the resolution of Maxwell's
demon paradox \cite{bennett,penrose,md2}, an important factor that
stimulates continued interest on the thermodynamics of information
processing. The demon has also been investigated quantum
mechanically \cite{zurek,lloyd,bender,md2}. Also, several
alternative proofs \cite{proof1,proof2,jacobs,yu} and
generalizations \cite{maroney-abs,maroney-gen,turgut} of LEP have
been given and different approaches aimed at quantum information
processing have been developed \cite{anderson,sagawa}. The crucial
ingredient in all alternative approaches to LEP is the description
of the operation on the device as a unitary transformation on a
larger combined system of the device and the bath where initially
the two are uncorrelated and the bath is in thermal
equilibrium\footnote{For classical devices, the analogous feature
of the transformation on the combined system is the fact that this
is a canonical map, which preserves phase-space volumes by
Liouville's theorem. LEP can be deduced by following a similar
line of reasoning.}. Unitarity of this transformation is an
absolute necessity for all physical operations \cite{oqs} while
the noncorrelation and equilibrium assumptions are well-suited for
all possible experiments. Although this viewpoint is contested
\cite{norton}, LEP directly follows from the unitarity: in a
Landauer erasure, the transformation maps different initial
device-states to the same final device-state. Since the unitary
transformation has to be one-to-one, this can only happen if the
state of the bath is also modified. Finally, as the bath was
initially in equilibrium, changes in its state imply an increase
in its energy. Exact calculations then show that the average
increase in the energy of the bath must exceed the Landauer bound.
Can there be further implications of the unitarity of the transformation
on the energy exchange with the bath? A recent article gives an affirmative answer
to this question for
the case of processes that manipulate classical information \cite{turgut}. That article presents an
alternative formulation of LEP in terms of the detailed heat
transfers into the bath. In this non-entropic approach, one takes
the individual heats dumped to the bath for each initial logical
state as the main quantities of interest and places restrictions
between them. For example, if $q_i$ is the heat dumped to the bath
during a Landauer erasure when the initial logical state is $i$,
then it can be shown that
\begin{equation}
\sum_i e^{-q_i/k_B T} \leq 1\quad,
\label{eq:classical_landauer_exp}
\end{equation}
an inequality which is first derived by Szilard \cite{szilard}.
The main advantage of this formulation is its independence of the
coding of the information into the logical states of the device.
The standard statement of LEP, however, is dependent on the
coding: if the device is used for storing information in such a
way that the logical state $i$ is used with frequency $p_i$, then
the device has information storage capacity of $I=-\sum_i
p_i\log_2 p_i$ bits and the Landauer bound states that
\begin{equation}
\bar{q} \geq Ik_BT\ln2
\label{eq:classical_landauer_avg}
\end{equation}
where $\bar{q}=\sum_i p_i q_i$ is the average heat dumped to the
bath when the information in the device is erased.
Note that Eq.~(\ref{eq:classical_landauer_avg}) is valid for any
distribution $p$, hence it provides an infinite number of
restrictions on $q_i$. It can be shown that
Eq.~(\ref{eq:classical_landauer_avg}) holds for all possible
distributions $p$ if and only if
Eq.~(\ref{eq:classical_landauer_exp}) is satisfied. Thus,
Eqs.~(\ref{eq:classical_landauer_exp}) and
(\ref{eq:classical_landauer_avg}) are two equivalent formulations
of LEP. The standard form of LEP,
Eq.~(\ref{eq:classical_landauer_avg}) is more profound as it
directly connects two disparate disciplines, thermodynamics and
information theory. Yet, Eq.~(\ref{eq:classical_landauer_exp}) is
also somewhat attractive as it is a concise inequality that
captures all of the infinitely many inequalities in
Eq.~(\ref{eq:classical_landauer_avg}). Both inequalities can be
generalized to arbitrary nondeterministic processes. It is
observed that the generalization of
Eq.~(\ref{eq:classical_landauer_exp}) contains further
restrictions on heat exchanges that cannot be derived from the
entropic restrictions of the generalized LEP \cite{turgut}.
The purpose of this article is to adapt that coding-independent
approach to quantum operations applied on devices holding quantum
information. The major difference between the classical and
quantum cases is in the description of the state: instead of a
discrete index $i$, we now have a density matrix that describes
the state of the device. Hence, in place of the numbers $q_i$, we
now have a hermitian operator $Q$ whose expectation value with the
density matrix gives the average heat dumped to the bath. This
operator will be called as the \emph{heat transfer operator}
(HTO).
The unitary transformation
can be thought as moving information around
in the combined system of the device and the bath. The resulting
information exchange between the device and the bath changes the
state and hence the energy of the bath. The HTO essentially
provides a measure of this information exchange in terms of the
energy given to the bath. The nature of the operation applied on
the device (e.g., whether the information in the device is erased
completely or only partially, or whether randomness from the bath
is introduced to the device or not) has a strong effect on the
HTO. This article investigates that relation between quantum
operations and the associated HTOs. This relation essentially
follows from the unitarity of the transformation on the combined
system. However, it appears that the unitarity property is
unnecessarily too restrictive; all of the results in this article
can also be obtained from a weaker assumption, namely that the
transformation is an isometry. This assumption is made throughout
the article.
The quantum operation-HTO relation is a generalization of the
classical LEP to the quantum domain and hence it is primarily of
theoretical interest. But, it is possible that this may find some
applications as well. For example, the restrictions on HTOs can be
used as a way to theoretically quantify the ``information erasing
nature'' or some other aspect of generic quantum operations. In
addition to this, the current work can also be useful in the
design of non-unitary operations (e.g., erasures) on technological
implementations of quantum devices. If the heat emitted during the
operation is desired to be minimized, it is necessary to select or
adjust the bath levels and the device-bath interaction in a
suitable way. The explicit construction of the erasure operations
that appear in this article is appropriate for that optimization
task and hence it may be a valuable guide for this purpose.
The organization of the article is as follows. In
Sec.~\ref{sec:hto}, the HTO associated with a realization of a
given quantum operation is defined. The nature of the HTO and some
of its fundamental properties are also discussed in this section.
Sec.~\ref{sec:Restrictions} is devoted to the derivation of
various restrictions on the HTOs. The first one of these is the
entropic bound of the generalized LEP, which can be expressed for
an arbitrary quantum operation. This bound is a good starting
point for the investigation of the coding-independent
characterization of quantum operation-HTO relation. After that,
quantum operations that completely erase the information in the
device are taken up and the associated HTOs are described. These
results are then used for the description of HTOs associated with
extremal quantum operations. In Sec.~\ref{sec:differences}, some
examples are given for quantum operations that do not completely
erase the information content of the device and it is shown that
LEP is not sufficient for a complete characterization of HTOs
associated with these operations. This indicates the usefulness of
the current approach. Finally, Sec.~\ref{sec:conclusions} contains
brief conclusions.
\section{Heat Transfer Operator}
\label{sec:hto}
Let A be a quantum system with a finite number of levels which is
used as a memory device for holding quantum information. A quantum
operation on this system is a trace-preserving completely-positive
map $\mathcal{E}$ that takes the initial state $\rho_A$ of the
device to a final state $\mathcal{E}(\rho_A)$. Although all
quantum operations are within the scope of this article, special
attention will be given to those that completely erase all
information in the device. The operation $\mathcal{E}$ will be
called a \emph{complete erasure} if the final state is a constant
mixed state $\rho_0$ for all initial states, i.e.,
\begin{equation}
\mathcal{E}(\rho)=\rho_0\quad\textrm{for~all~}\rho~~.
\end{equation}
It will be called a \emph{Landauer erasure} if it is a complete
erasure with a pure final state, i.e.,
$\rho_0=\ket{\psi_0}\bra{\psi_0}$ for some $\ket{\psi_0}$.
Any quantum operation $\mathcal{E}$ on A can be realized as a unitary
transformation or an isometry $U_{AB}$ on a larger system AB,
where B denotes an additional system that
will henceforth be called as the bath. It is assumed that the device and
the bath are initially uncorrelated. Hence, if the initial state of the
bath is $\rho_B$, the initial state of the composite system is the product
state $\rho_{AB}=\rho_A\otimes\rho_B$. The final states after the
application of the isometry will usually be denoted by primes, e.g.,
$\rho_{AB}^\prime=U_{AB}(\rho_A\otimes \rho_B) U_{AB}^\dagger$. We will
say that the bath B, the state $\rho_B$ and the isometry $U_{AB}$ form a
\emph{realization} of the operation $\mathcal{E}$ if
\begin{equation}
\mathcal{E}(\rho_A)=\rho_A^\prime=\tr_B \left( U_{AB}(\rho_A\otimes \rho_B)U_{AB}^\dagger\right)
\label{eq:realization_def}
\end{equation}
is satisfied for all $\rho_A$, where here $\tr_B$ denotes the
partial trace over B.
It is assumed that the transformation $U_{AB}$ is obtained from
the evolution of the composite system AB with a time-dependent
Hamiltonian $H_{AB}(t)$. The basic assumption about the
independence of the process (i.e., the Hamiltonian and therefore
the transformation $U_{AB}$) from the information stored in the
device is also made. In particular this means that no measurements
are taken on the composite system AB during the process. Hence,
during the transformation, any information exchange with A occurs
entirely between A and B only and this process is described with
the Hamiltonian $H_{AB}(t)$. No third system is involved in the
transformation except in providing the energy for the necessary
work done on AB.
For discussing the energetics of the aforementioned
transformation, the time-dependent Hamiltonian $H_{AB}(t)$ of the
composite system AB has to be discussed. The transformation
process is assumed to be carried out over a finite time interval,
say between times $t_i$ and $t_f$, during which subsystems A and B
couple and $H_{AB}(t)$ is time dependent. Outside this time
interval, however, it is assumed that the two subsystems are
uncoupled and have time-independent Hamiltonians. Thus, we have
\begin{eqnarray}
H_{AB}(t) &=& H_A + H_B \quad\textrm{for}~t < t_i~\textrm{and} \label{eq:Hamiltonian_before} \\
H_{AB}(t) &=& H_A^\prime + H_B \quad\textrm{for}~t > t_f~, \label{eq:Hamiltonian_after}
\end{eqnarray}
where $H_A$ and $H_A^\prime$ are the Hamiltonians of the device
before and after the transformation respectively and $H_B$ is the
bath Hamiltonian. The Hamiltonian of the device is not relevant to
the main subject of the current article and hence there are no
restrictions on $H_A$ and $H_A^\prime$. However, initial and final
Hamiltonians of the bath are taken to be identical. This is also
operationally reasonable for realistic quantum operations where
the environment corresponds to B; a quantum operation on a device
may perhaps structurally change the device (so that initial and
final Hamiltonians may differ) but it merely couples with the
environment without changing it structurally. Finally, no
assumptions are made about the strength of the coupling and the
Hamiltonian $H_{AB}(t)$ during the time interval $t_i\leq t\leq t_f$.
The transformation $U_{AB}$ is the time-development operator
corresponding to the Hamiltonian $H_{AB}(t)$ between two times
$t_1$ and $t_2$ where $t_1\leq t_i$ and $t_f\leq t_2$. As
indicated in Eqs.~(\ref{eq:Hamiltonian_before}) and
(\ref{eq:Hamiltonian_after}), it is assumed that A and B do not
interact with each other outside the process time interval,
$(t_i,t_f)$. For this reason, no special values are needed to be
assumed about $t_{1,2}$. However, the results of the current
article can also be extended to cases where there is a weak
coupling between A and B outside the process time interval. In
this case, $t_2-t_f$ should be greater than the
thermal-equilibration time scales of the bath.
Note that for all realistic processes the operator $U_{AB}$ is
unitary, i.e., it satisfies $U_{AB}^\dagger
U_{AB}=U_{AB}U_{AB}^\dagger=\mathds{1}_{AB}$. However, for the
derivation of the results in this article, this requirement
appears to be unnecessarily too restrictive; we only need to
assume that $U_{AB}$ is an isometry, i.e., only $U_{AB}^\dagger
U_{AB}=\mathds{1}_{AB}$ is assumed to be satisfied. Therefore, we
allow the possibility that the transformation $U_{AB}$ is not
surjective. The generalization of unitary transformations to
isometries also serves another purpose. There are some quantum
operations, like Landauer erasures, which can only be realized by
proper isometries instead of unitaries. This is quite obvious for
Landauer erasures since all possible accessible final states do
not span the whole of the Hilbert space of AB, because any initial
pure state of AB is mapped to a final product state where A is in
$\ket{\psi_0}_A$ state\footnote{In this work, it is required that
the bath is initially in a canonical equilibrium state. Therefore
all pure states of B are possible as initial states and the map
$U_{AB}$ should send all of these to a state where A is in
$\ket{\psi_0}$ state.}.
As $U_{AB}$ is not surjective, it cannot possibly be unitary. Such
proper isometries can in principle be realized by a Hamiltonian
evolution, provided that ideal tools like perfectly impenetrable
barriers are used for making parts of the Hilbert space
inaccessible.
It is assumed that the bath is in canonical thermal
equilibrium state before the transformation and thus
\begin{equation}
\rho_B = \frac{1}{Z_B}\exp\left(-\beta H_B\right)\quad,
\end{equation}
where $\beta=1/k_B T$ is the inverse temperature and $Z_B$ is the
corresponding partition function. It is appropriate to also assume
that the bath is sufficiently small so that $\rho_B$ is
manageable. We will say that B is a \emph{finite} system if $Z_B$
is finite for all temperatures ($Z_B\neq\infty$). In this case, B
has finite thermodynamical properties. Note that a single Hydrogen
atom or a single free particle placed in an infinite space has an
infinite partition function and therefore these are not finite
systems by the current definition. But, if the atom or the
particle is placed inside a large but finite box, then they will
become finite systems. Note also that any realistic system, which
has a finite number of particles constrained inside a finite
volume, will be classified as finite by this definition. Infinite
systems have individually vanishing level occupation probabilities
and consequently the analysis of such systems present some
difficulties. This is the main reason for using finite systems in
this article.
Let $E_B=\tr\,\rho_BH_B$ be the equilibrium energy of the bath.
Since there is no structural change in the bath, the heat dumped
to the bath during the realization is equal to the change in the
average energy, $\Delta E_B$, between times $t_1$ and $t_2$. It
can be expressed as
\begin{eqnarray}
\Delta E_B &=& \tr (\rho_B^\prime - \rho_B) H_B \quad,
\label{eq:Delta_EB_defined} \\
&=& \tr_{AB} \left(\rho_A\otimes \rho_B\right) \left(U_{AB}^\dagger(\mathds{1}_A\otimes H_B )U_{AB}\right)
-E_B \nonumber
\\
&=& \tr_A (\rho_A Q)\nonumber
\end{eqnarray}
where $Q$ is the HTO of this realization, which is defined as
\begin{equation}
Q = \tr_B \left[ \left(\mathds{1}_A\otimes \rho_B\right) U_{AB}^\dagger(\mathds{1}_A\otimes H_B )U_{AB} \right]
- E_B \mathds{1}_A ~.
\label{eq:Q_definition}
\end{equation}
A number of comments have to be made in order to comprehend the
real nature of the HTO. First, $Q$ is a hermitian operator defined
on the state space $\mathcal{H}_A$ of the device A. Its
expectation value is evaluated with the initial state, $\rho_A$,
of the device. In other words, it is an operator that can be used
for expressing the energy transferred to B in terms of the state
of A. Second, only the average of $Q$ can be given a physical
meaning. It has been shown that no quantum observable can be
defined for the work done or the heat transferred during a process
that captures the detailed distribution of these
quantities\cite{hanggi}. However, if only the average $\Delta E_B$
is needed, then Eq.~(\ref{eq:Delta_EB_defined}) is the correct
definition, which eventually yields the HTO $Q$. On the other
hand, $Q$ cannot be used for computing the higher moments of the
energy transferred to B. For example, even though the average
energy transferred to B is given by $\tr\,\rho_AQ$, the quantity
$\tr\,\rho_AQ^2$ or the averages of other nonlinear functions of
$Q$ do not have the anticipated meaning. Averages of such
nonlinear functions of energy change of B should be computed
separately following the approach in Ref.~\onlinecite{hanggi}.
This crucial property of the HTO must always be kept in mind.
Once the HTO is computed, the average total work that needs to be
done for carrying out the operation can be obtained as
\begin{equation}
W_{\textrm{tot}}= \tr\left(\mathcal{E}(\rho_A) H_A^\prime -\rho_A H_A+\rho_A Q\right)~.
\end{equation}
As a result, for investigating the average energies transferred
during the realization of the operation $\mathcal{E}$, the only
nontrivial quantity to be determined is the heat dumped to the
bath. This is the main reason why this article concentrates solely
on the HTO.
The isometry condition $U_{AB}^\dagger U_{AB}=\mathds{1}_{AB}$ and
the fact that the bath is in thermal equilibrium imposes some
restrictions on the operator $Q$. The investigation of these
restrictions is the primary subject of this article. The main
question that needs to be answered is the following: for any given
quantum operation $\mathcal{E}$, hermitian operator $Q$ and
temperature value $T$, is it possible to find a bath B (namely a
Hamiltonian $H_B$) and an isometry $U_{AB}$ such that both
Eq.~(\ref{eq:realization_def}) and Eq.~(\ref{eq:Q_definition}) are
satisfied? We will say that $Q$ is a possible HTO for
$\mathcal{E}$ when such a bath and isometry exist. Our main
problem is then to find all necessary and sufficient conditions
for $Q$ to be a possible HTO for $\mathcal{E}$ at temperature $T$.
These conditions will enable us to discuss the thermodynamics of
quantum information processing (for given $\mathcal{E}$) without
specifying anything about the bath, the bath-device coupling and
the specific way the quantum operation is realized.
Before embarking on the investigation of this problem, it will be
convenient to list the following basic properties of the HTOs.
(i) First, note that any restriction that the HTOs obey can be
expressed entirely in terms of $\mathcal{E}$ and the ratio
$Q/k_BT=\beta Q$. To see this, note that both $\mathcal{E}$ and
$\beta Q$ can be expressed exclusively in terms of $U_{AB}$ and
$\rho_B$, with no explicit $T$ (or $H_B$) dependence, i.e.,
Eq.~(\ref{eq:realization_def}) and
\begin{eqnarray}
\beta Q &=& - \tr_B \left[ \left(\mathds{1}_A\otimes \rho_B\right) U_{AB}^\dagger(\mathds{1}_A\otimes \ln\rho_B )U_{AB} \right]
\nonumber \\
& & \quad + \mathds{1}_A\tr_B(\rho_B\ln\rho_B)\quad.
\label{eq:Q_definition_alternative}
\end{eqnarray}
Hence, the main problem becomes: for given $\mathcal{E}$ and
$\beta Q$, can we find an appropriate $\rho_B$ and $U_{AB}$ such
that Eqs.~(\ref{eq:realization_def}) and
(\ref{eq:Q_definition_alternative}) are satisfied?
(ii) If $Q$ is a possible HTO for an operation $\mathcal{E}$,
then for any positive number $a$, the operator
$\tilde{Q}=Q+a\mathds{1}_A$ is also a possible HTO for the same
operation. In other words, it is always possible to dump an
additional heat $a$ whose amount is independent of the state of the device.
To show this, suppose that $U_{AB}$ on bath B is a realization of
$\mathcal{E}$ that yields the HTO $Q$. Let C be a two-level
system with Hamiltonian $H_C=\Delta \sigma_z/2$ where $\sigma_z$ denotes the associated
Pauli spin matrix. Now, take BC as the new bath and apply the isometry
$\tilde{U}_{ABC}=U_{AB}\otimes (\sigma_x)_C$ to the composite system ABC.
Since the transformation that C is subjected to is independent of how AB
is changed, it is obvious that $\tilde{U}_{ABC}$ realizes the same operation $\mathcal{E}$.
But, the heat dumped to the bath BC is now
\begin{equation}
\Delta E_{BC} = \tr (\rho_AQ)+\Delta E_C\quad,
\end{equation}
where $\Delta E_C=\Delta\tanh(\Delta/2k_BT)$. If the value of $\Delta$ is
chosen such that $\Delta E_C=a$, then we get $\Delta
E_{BC}=\tr\rho_A\tilde{Q}$. This shows that $\tilde{Q}$ is also a possible
HTO for the same operation.
(iii) The set of two-tuples $(\mathcal{E},Q)$ formed from quantum
operations $\mathcal{E}$ and allowed HTOs $Q$ is a convex set. In other
words, if $Q_i$ is a possible HTO for the operation $\mathcal{E}_i$ for
$i=1,2,\ldots,n$, then for any $\lambda_i\geq0$ with $\sum_i\lambda_i=1$,
the operator $Q=\sum_i\lambda_iQ_i$ is a possible HTO for the operation
$\mathcal{E}=\sum_i\lambda_i\mathcal{E}_i$.
Before showing this, first note that realizations for each $\mathcal{E}_i$
can be assumed to involve the same bath B. This can always be done by
taking the bath sufficiently large. As a result, without loss of
generality, suppose that $U_{AB}^{(i)}$ is an isometry on AB that realizes
$\mathcal{E}_i$ and yields the associated HTO $Q_i$. Let C be an $n$-level
quantum system and let $\{\ket{1},\cdots,\ket{n}\}$ be an orthonormal
basis in the Hilbert space of C. Let the initial state of C be
$\rho_C=\sum_i \lambda_i \ket{i}\bra{i}$. Now, consider the isometry
$V_{ABC}$ on ABC given by
\begin{equation}
V_{ABC}=\sum_{i=1}^n U_{AB}^{(i)} \otimes \left(\ket{i}\bra{i}\right)_C
\end{equation}
It is a simple exercise to show that this isometry gives rise to
the desired convex combinations, i.e., $V_{ABC}$ realizes the
operation $\mathcal{E}=\sum_i\lambda_i\mathcal{E}_i$ and the
associated HTO is $Q=\sum_i\lambda_iQ_i$. Note that, in here, C
acts like a \emph{controller} system for deciding which operation
should be applied on AB. The decision is given probabilistically
such that $\mathcal{E}_i$ is applied with probability $\lambda_i$.
Note also that the average energy of C does not change during this
process. In such a case, the overall operation and the associated
HTO are given by the expected convex combination expressions.
(iv) As a special case of the convexity result in (iii), the set of
allowed HTOs associated with a fixed quantum operation $\mathcal{E}$ is
also convex. This set is not bounded from above by (ii); but it is bounded
from below. We are primarily interested in finding lower bounds, if
possible tight ones, for this set.
\section{Restrictions on Heat Transfer Operators}
\label{sec:Restrictions}
\subsection{Entropic Restriction of LEP}
Before investigating the relations between HTOs and quantum
operations, it is appropriate to start with the state-dependent
entropic restriction of LEP. Let $S(\rho)=-\tr\,\rho\ln\rho$
denote the von Neumann entropy of the density matrix $\rho$. For
an arbitrary quantum operation, the average heat emitted to the
bath is bounded from below by the drop in the von Neumann entropy
of the state of the device.
\begin{theorem}[Generalized LEP]
\label{thm:genlep}
If $Q$ is a possible HTO for an operation $\mathcal{E}$, then
\begin{equation}
\tr\,\rho_A Q \geq k_B T \Big[S(\rho_A)-S(\mathcal{E}(\rho_A))\Big]\quad,
\label{eq:entropic_LEP}
\end{equation}
for all states $\rho_A$ of the device A. For a given $\rho_A$, the
equality sign in (\ref{eq:entropic_LEP}) holds if and only if the
relation $\rho_{AB}^\prime = \mathcal{E}(\rho_A)\otimes \rho_B$ is
satisfied for some (and therefore for all) realization that yields
$\mathcal{E}$ and $Q$.
\end{theorem}
Note that the inequality above essentially gives infinitely many
bounds on the HTO, i.e., there is a separate bound on $Q$ for each
possible device state $\rho_A$. The generalized LEP has been
discussed in several
works\cite{landauer,penrose,maroney-abs,maroney-gen,turgut},
especially in the context of classical information processing.
However, for the purpose of treating the equality case, the
following straightforward derivation is included in here. It can
be deduced simply from the following wonderful identity for
deviations from canonical thermal equilibrium states
\begin{equation}
\Delta E_B = k_B T \left( S(\rho_B^\prime\vert\vert\rho_B) +S(\rho_B^\prime)-S(\rho_B)\right)
\label{eq:magic}
\end{equation}
where
\begin{equation}
S(\sigma\vert\vert\rho)=\tr\,\sigma(\ln\sigma-\ln\rho)
\end{equation}
denotes the relative entropy function. Only two inequalities will
be used for converting this identity to the relation in
Eq.~(\ref{eq:entropic_LEP}). The first one is the nonnegativity of
the relative entropy: $S(\sigma\vert\vert \rho)\geq 0$ with
equality holding if and only if $\sigma=\rho$. The second
inequality is the subadditivity of the von Neumann entropy for the
final state, i.e., $S(\rho_{AB}^\prime)\leq
S(\rho_A^\prime)+S(\rho_B^\prime)$ where the equality holds if and
only if the state is in product form
$\rho_{AB}^\prime=\rho_A^\prime\otimes\rho_B^\prime$. Finally,
using the fact that the initial and final states are related by an
isometry we conclude that $\rho_{AB}^\prime$ and $\rho_{AB}$ are
isospectral. In particular,
$S(\rho_{AB}^\prime)=S(\rho_{AB})=S(\rho_A)+S(\rho_B)$ and hence
\begin{equation}
\Delta E_B\geq k_BT ( S(\rho_B^\prime)-S(\rho_B))\geq k_BT(S(\rho_A)-S(\rho_A^\prime)) ~,
\nonumber
\end{equation}
which is the desired inequality. It is easy to see that both of
the inequalities become equalities if and only if
$\rho_{AB}^\prime = \mathcal{E}(\rho_A)\otimes \rho_B$.$\Box$
Note that the equality can be satisfied only in some exceptional cases.
When that happens, we have $\rho_{AB}^\prime = \mathcal{E}(\rho_A)\otimes
\rho_B$ and therefore the state of the bath is unchanged, i.e.,
$\rho_B^\prime=\rho_B$. Consequently, the average heat dumped to bath is
$\tr\,\rho_A Q=0$. Moreover, since the initial and final density matrices
of AB are isospectral, it follows that $\rho_A$ and $\mathcal{E}(\rho_A)$
must also be isospectral.
It is of some interest to investigate the issue of whether the
bound in (\ref{eq:entropic_LEP}) can be improved by considering
the device A to be part of a larger system AX where X is not
involved in the realization. In other words, the inequality
\begin{equation}
\tr\,\rho_A Q \geq k_BT\left(S(\rho_{AX})-S((\mathcal{E}\otimes\mathcal{I})(\rho_{AX}))\right)
\label{eq:entropic_LEP_X}
\end{equation}
is also valid for any X and any initial state $\rho_{AX}$. Can this
generalized inequality be tighter than Eq.~(\ref{eq:entropic_LEP})? The
answer is negative. To show this, introduce another system Y such that AXY
is a purification of $\rho_{AX}$. The isometry of the realization is
applied on AB and hence the system XY is untouched. Using primes for
denoting the final states we have
$S(\rho_A)=S(\rho_{XY})=S(\rho_{XY}^\prime)$ and
$S(\rho_{AX})=S(\rho_Y)=S(\rho_Y^\prime)$. Then we invoke the following
alternative form of the strong subadditivity of the von Neumann entropy
\cite{Nielsen}
\begin{equation}
S(\rho_{XY}^\prime)+S(\rho_{XA}^\prime) \geq S(\rho_Y^\prime) +
S(\rho_A^\prime)
\end{equation}
to arrive at
\begin{equation}
S(\rho_A) - S(\rho_{A}^\prime) \geq S(\rho_{AX}) - S(\rho_{AX}^\prime)\quad.
\end{equation}
The last inequality clearly shows that the lower bound given by
Eq.~(\ref{eq:entropic_LEP_X}) is always smaller than the bound in
(\ref{eq:entropic_LEP}). Hence, the inclusion of additional
systems cannot possibly make an improvement in the entropic bounds
of LEP.
\subsection{Restrictions for Complete Erasure Operations}
Even though Theorem \ref{thm:genlep} places some restrictions on the
possible HTOs, it is possible that these may not be complete. In other
words, there might be operators $Q$ which satisfy the inequality
(\ref{eq:entropic_LEP}) for all possible initial states $\rho_A$, but it
still may not be a possible HTO. Some examples of such operators will be
discussed in Sec.~\ref{sec:conclusions}. However, it appears that the
inequalities (\ref{eq:entropic_LEP}) are sufficient when $\mathcal{E}$ is
a complete erasure operation. This subsection contains the proof of this
statement.
But, before that, a concise form of the inequalities in
(\ref{eq:entropic_LEP}) should be found and for this reason it is
necessary to introduce a real-valued function of hermitian
operators. The needed function is the Legendre transform $J$ of
the von Neumann entropy function $S$. For any hermitian $G$, this
function is defined by
\begin{equation}
J(G) = -\ln \left(\tr\, e^{-G}\right)\quad.
\end{equation}
In order to arrive at the main property of this function, first
note that for any hermitian operator $G$, the operator
\begin{equation}
\sigma_G =\exp(J(G))e^{-G}
\end{equation}
is a density matrix. Then, using the nonnegativity of
$S(\rho\vert\vert\sigma_G)$ we find
\begin{equation}
J(G) + S(\rho) \leq \tr\,\rho G
\end{equation}
an inequality which is valid for any density matrix $\rho$ and any
hermitian $G$. From this inequality it is possible to derive the
following Legendre transformation identities
\begin{eqnarray}
J(G) &=& \min_\rho \left(\tr\, \rho G - S(\rho)\right)\quad, \label{eq:Legendre_J_from_S}\\
S(\rho) &=& \inf_G \left(\tr\, \rho G - J(G) \right)\quad, \label{eq:Legendre_S_from_J}
\end{eqnarray}
where in Eq.~(\ref{eq:Legendre_J_from_S}) the minimization is over
all density matrices and the minimum is reached when
$\rho=\sigma_G$. In Eq.~(\ref{eq:Legendre_S_from_J}) the infimum
is over all hermitian operators $G$. If $\rho$ has no zero
eigenvalues, then right-hand side of that equation is minimized by
$G=-\ln \rho$. Otherwise, if $\rho$ has zero eigenvalues, then the
right-hand side of (\ref{eq:Legendre_S_from_J}) has no minimum;
but infimum is obtained by a sequence of density matrices which
approach to the limit $G_{\textrm{lim}}=-\ln\rho$, which is now an
operator that has some infinite eigenvalues.
Consider a complete erasure operation $\mathcal{E}$ where
$\mathcal{E}(\rho_A)=\rho_0$ for all states $\rho_A$. In this
case, the inequality (\ref{eq:entropic_LEP}), which is valid for
all density matrices $\rho_A$, can be written compactly using the
$J$-function as
\begin{equation}
J(\beta Q) \geq -S(\rho_0)\quad.
\label{eq:Cond_Complete_Erasure}
\end{equation}
Moreover, this condition is almost sufficient for $Q$ to be a
possible HTO. There is only a slight complication in the treatment
of the equality case which is related to the finiteness property
of the bath. The following theorem gives the complete
characterization of HTOs for complete erasures.
\begin{theorem}
\label{thm:complete} Let $Q$ be a hermitian operator and $\mathcal{E}$ be
a complete erasure to the final state $\rho_0$. When only realizations
involving finite baths are considered, $Q$ is a possible HTO for
$\mathcal{E}$ if and only if
\begin{itemize}
\item[(i)] either $J(\beta Q)>-S(\rho_0)$,
\item[(ii)] or $J(\beta Q)=-S(\rho_0)$ and $\sigma_{\beta Q}$ is
isospectral with $\rho_0$.
\end{itemize}
\end{theorem}
Therefore the equality sign in (\ref{eq:Cond_Complete_Erasure})
applies only under very rare situations where the eigenvalue
spectrum of $Q$ must satisfy strict conditions; in particular the
dimension of the state space of A must be equal to the matrix rank
of the final state $\rho_0$. If this is not the case, then the
strict inequality holds.
The proof of the necessity of the conditions (i) and (ii) has
already been argued above in obtaining
Eq.~(\ref{eq:Cond_Complete_Erasure}); only a treatment of the
equality case is left. If equality holds in
(\ref{eq:Cond_Complete_Erasure}), then the density matrix that
minimizes the expression (\ref{eq:Legendre_J_from_S}), namely
$\rho_A=\sigma_{\beta Q}$, gives equality in
(\ref{eq:entropic_LEP}), and therefore by Theorem
\ref{thm:genlep}, $\rho_A$ must be isospectral with
$\mathcal{E}(\rho_A)=\rho_0$, i.e., the condition in (ii) holds.
This completes the proof of the necessity of the conditions of
Theorem \ref{thm:complete}.
For the sufficiency proof, we have to start with a given $Q$ that
satisfies either (i) and (ii) and then show that a realization
that yields $\mathcal{E}$ and $Q$ can be constructed. The needed
isometry is actually quite simple. It essentially copies the
entire information on the device to some subsystem of the bath, so
that the information is completely erased, and at the same time it
copies the thermal equilibrium state of some subsystem of B to the
device, so that the device has always the same final state. Even
though the isometry is simple, the job of adjusting the HTO to be
equal to the given operator $Q$ is technically complicated. For
this reason, the proof of the sufficiency of the conditions in
Theorem \ref{thm:complete} is done separately in Appendix
\ref{sec:app_com}. However, it is appropriate to give a separate
proof in here for the special case of Landauer erasures for which
the required bath is simpler. Treatment of the technical details
in this proof will also simplify the presentation of the full
proof of Theorem \ref{thm:complete} in Appendix \ref{sec:app_com}.
Let $\mathcal{E}$ be a Landauer erasure to the standard state
$\ket{\psi_0}$ ($\mathcal{E}(\rho_A)=\ket{\psi_0}\bra{\psi_0}$ for
all $\rho_A$). Note that the special case (ii) of Theorem
\ref{thm:complete} can hold only if the dimension of the Hilbert
space of A is $d=1$. This is a very trivial case where the system
A cannot hold any information at all. In such a case, the HTO $Q$
is a number and the theorem just says that this number is
nonnegative, $Q\geq0$. For the generic case of $d\geq2$, however,
only the strict inequality of (i) holds. The following is merely a
corollary of Theorem \ref{thm:complete} but we state it as a
separate theorem and give a separate proof.
\begin{theorem}
\label{cor:Landauer}
Let $\mathcal{E}$ be a Landauer erasure on a device A having
$d$ levels with $d\geq2$ and let $Q$ be a hermitian operator. The
following are equivalent.
\begin{itemize}
\item[(i)] $Q$ is a possible HTO for a realization on a finite
bath.
\item[(ii)] $\tr\, e^{-\beta Q} < 1$.
\item[(iii)] $\tr\, \rho_A Q>k_BT S(\rho_A)$ for all $\rho_A$.
\end{itemize}
\end{theorem}
It is obvious that the statements (ii) and (iii) are equivalent;
both are simply stating that $J(\beta Q)$ is a strictly positive
number. The statement (iii) is the conventional way of expressing
the LEP. Statement (ii) is the quantum analog of the inequality in
Eq.~(\ref{eq:classical_landauer_exp}). (The strict inequality is a
minor mathematical detail; it arises from our insistence on using
finite baths.)
It has been shown in Theorem \ref{thm:genlep} that (i) implies (ii) and
(iii). What is left is the proof of the implication (ii)$\Rightarrow$(i).
Let $Q$ be a hermitian operator such that $J(\beta Q)>0$. A realization of
the Landauer erasure will be constructed below and it will be shown that
the associated HTO is equal to $Q$.
\begin{figure}
\includegraphics[scale=0.8]{fignew1.eps}
\caption{The pictorial description of the isometry used in the
proof of Theorem \ref{thm:complete} for a Landauer erasure
operation. Here, the state of the device A is made equal to the
standard state $\ket{\psi_0}$ while the original information
content of each subsystem is shifted to the right.}
\label{fig:Landauer}
\end{figure}
\paragraph{The bath and its basis states.}
It is obvious that the bath needed for this purpose has infinitely
many levels. Only in that case, it is possible to copy the
information on A onto B and at the same time, retain the
information already present in B. For better visualizing the
information flow, we will think of the bath as being composed of
infinitely many copies of a $d$-level system X. The constitution
of B can be expressed as
$\mathrm{B}=\mathrm{X}_1\mathrm{X}_2\mathrm{X}_3\cdots$. Let the
standard orthonormal basis of X be denoted by $\{\ket{i}\}$ where
$i=0,1,\ldots,d-1$. The following self-explanatory shorthand will
be used for the state vectors of the bath
\begin{equation}
\ket{i_1,i_2,i_3,\cdots}_B = \ket{i_1}_{X_1}\otimes\ket{i_2}_{X_2}\otimes\ket{i_3}_{X_3}\otimes\cdots~~.
\end{equation}
The Hilbert space $\mathcal{H}_B$ of the bath will be defined as
the linear span of the above states where the sequence $\{i_k\}$
contains only \emph{finitely many nonzero elements}. When the
Hamiltonian is defined below, it will be seen that the states
which are not included into $\mathcal{H}_B$ have infinite energy
and zero equilibrium-state probabilities. Consequently such states
will never enter into the discussion.
In addition to this, defining $\mathcal{H}_B$ in such a way makes
it a \emph{separable} Hilbert space having a countable orthonormal
basis. To see this, note that the sequence $\{i_k\}$ can be
considered as the digits of the base-$d$ representation of an
integer $n$, written in short as $n=(\cdots i_3i_2i_1)_d$. The
integer $n$ is given in terms of the sequence as
\begin{equation}
n = i_1+i_2 d+i_3d^2+\cdots=\sum_{k=1}^\infty i_k d^{k-1}~~.
\end{equation}
Consequently, there is a one-to-one correspondence between
nonnegative integers $n$ and sequences $\{i_k\}$ that has finitely
many nonzero elements. For this reason, the standard orthonormal
basis of $\mathcal{H}_B$ can be enumerated by a nonnegative
integer $n$ and therefore the alternative label
\begin{equation}
\ket{n}_B=\ket{i_1,i_2,i_3,\cdots}_B
\end{equation}
can be used. It is thus clear that the Hilbert space
$\mathcal{H}_B$ has a countable basis and hence it is separable.
\paragraph{The isometry.}
Let $q_i$ ($i=0,1,\ldots,d-1$) be the eigenvalues of $Q$ arranged
in nondecreasing order and let $\{\ket{i}\}$ be the orthonormal
basis of $\mathcal{H}_A$ formed from the corresponding
eigenvectors.
\begin{equation}
Q = \sum_{i=0}^{d-1} q_i \ket{i}\bra{i} \qquad \, (q_0\leq q_1\leq \cdots\leq q_{d-1})~.
\label{eq:Q_spectral_decomp}
\end{equation}
The isometry $U_{AB}$ on the composite system AB is defined as
follows
\begin{equation}
U_{AB}\ket{\ell}_A \otimes \ket{i_1,i_2,i_3,\cdots}_B
= \ket{\psi_0}_A \otimes \ket{\ell, i_1,i_2,\cdots}_B~~,
\end{equation}
for any allowed values of input labels. In words, the isometry
sets the state of A to the standard state $\ket{\psi_0}$ and
copies its contents onto the subsystem X$_1$. It also forms an
infinite chain of information flow inside the bath as the contents
of X$_k$ are copied onto X$_{k+1}$ ($k\geq1$) for all $k$. The
flow of information is pictorially depicted in
Fig.~\ref{fig:Landauer}. Essentially the isometry shifts the
information content of all subsystems of AB onto the next
subsystem on the right.
This flow is highly reminiscent of the motion of the guests in
Hilbert's Hotel Infinity paradox. In that hotel, all of the
infinitely many rooms are occupied. But, it is still possible to
accommodate a new guest by moving everyone to the next room. The
isometry $U_{AB}$ essentially does the same for information. It
retains the whole information already stored in B while it
squeezes additional information coming from A into B. Our hotel,
however, has an important difference from Hilbert's hotel: as it
follows from LEP, there is a nonzero check-in cost. In other
words, while squeezing new information into B, the average energy
of B increases. To be able to compute that increase, the bath
Hamiltonian must be specified.
\paragraph{The Hamiltonian of the bath.}
The Hamiltonian will be constructed such that each subsystem of
the bath is independent and the energy eigenstates are identical
with the standard basis states. The precise energy eigenvalues of
the subsystem X$_1$ will be chosen to be identical with those of
$Q$ (plus a possible shift), i.e., $H_{X_1}=Q+c_1\mathds{1}_{X_1}$
in the standard basis where here, $c_1$ is some constant. The
energy eigenvalues of the remaining subsystems X$_2$, X$_3$,
$\ldots$ will only shift the HTO by a multiple of identity. With
this choice, the energy change of the bath is $\Delta
E_B=\tr(\rho_AQ)+c_2$ where $c_2$ is a constant that is
independent of $\rho_A$. The value of $c_2$ depends on the energy
levels of X$_2$, X$_3$, $\ldots$. By adjusting these levels
appropriately, it is possible set $c_2=0$. This shows that the HTO
is $Q$ and completes the proof. Since the details are complicated,
the adjustment of the energy levels of other subsystems is
discussed in Appendix~\ref{sec:app_lan}. $\Box$
The realization described in the proof above can be used for
finding conditions for minimizing the average heat emission during
a Landauer erasure operation. Suppose that A is used as a memory
device in an implementation of a quantum computer. Suppose also
that A is subjected to a Landauer erasure at ambient temperature
$T_{\mathrm{env}}$ in some steps of the operation of the computer.
Let $\rho_A^{\mathrm{av}}$ denote the average state of the device
before the erasures are applied. The average heat emitted per
erasure is then $\bar{q}=\tr\,\rho_A^{\mathrm{av}}Q$ where $Q$ is
the HTO. The problem is to minimize $\bar{q}$ under the
restriction that $Q$ satisfies the conditions mentioned in Theorem
\ref{cor:Landauer}. According to Theorem \ref{thm:genlep},
$\bar{q}$ is bounded below by
$k_BT_{\mathrm{env}}S(\rho_A^{\mathrm{av}})$. A simple inspection
shows us that this bound is attained by a small error by the HTO
\begin{equation}
Q = -k_BT_{\mathrm{env}} \ln \rho_A^{\mathrm{av}} + \epsilon\mathds{1}_A
\end{equation}
where $\epsilon$ is a small positive number. The construction in
the proof of Theorem \ref{cor:Landauer} can then be followed for
adjusting the energy levels of the bath, the detailed Hamiltonian
$H_{AB}(t)$ during device-bath interaction and the consequent
isometry. In a realistic situation where the environment serves as
the bath B, it is not possible to modify the energy levels of B at
will and the Landauer bound may not be attained; but the
construction above can still be useful. For example, a finite
chain X$_1$X$_2\cdots$X$_N$ can be used for connecting the device
and the environment. By adjusting the energy levels of the
subsystems in the chain and using the same information flow,
$\bar{q}$ can still be significantly decreased.
\subsection{General Quantum Operations}
A generic quantum operation causes only a partial erasure of the quantum
information stored by the device by forming an entanglement between the
device and the bath during the realization. Even though the inequality
(\ref{eq:entropic_LEP}) continues to be valid for the generic case, it
does not necessarily give a complete characterization of the possible HTOs
as it will be discussed in Sec. \ref{sec:conclusions}. This subsection
contains a few basic results about the HTOs of generic quantum operations
that surpass (\ref{eq:entropic_LEP}) in implications.
Let $\mathcal{E}$ be a quantum operation on the device A. Such an operation
can be written in the Kraus form
\begin{equation}
\mathcal{E}(\rho_A) = \sum_{i=1}^n M_i \rho_A M_i^\dagger
\label{eq:Kraus}
\end{equation}
where the Kraus operators $M_i$ satisfy
\begin{equation}
\sum_{i=1}^n M_i^\dagger M_i = \mathds{1}_A\quad.
\end{equation}
Kraus operators can be chosen in various different
ways\cite{Nielsen}. However, a representation that has the
smallest number $n$ of terms will be preferred in this article. In
this case, the set of Kraus operators $\{M_1,\ldots,M_n\}$ are
linearly independent.
Now, let $U_{AB}$ be the isometry for a realization of the
operation $\mathcal{E}$. This isometry can be written as
\begin{equation}
U_{AB} = \sum_{i=1}^n (M_i)_A \otimes (L_i)_B
\label{eq:U_AB_expansion}
\end{equation}
for some operators $L_i$ acting on the Hilbert space of the bath
B. This can be shown by completing the set of operators
$\{M_1,M_2,\ldots,M_n\}$ to a basis for the operator space by
adding new, linearly-independent operators
$M_{n+1},M_{n+2},\cdots$ to the list, and then applying the
expansion in Eq.~(\ref{eq:U_AB_expansion}) with a larger upper
limit of summation. As $U_{AB}$ is a realization of $\mathcal{E}$,
we have
\begin{eqnarray}
\mathcal{E}(\rho_A) &=& \tr_B U_{AB}(\rho_A\otimes \rho_B)U_{AB}^\dagger \\
&=& \sum_{ij} M_i \rho_A M_j^\dagger \left( \tr\,\rho_BL_j^\dagger L_i\right).
\end{eqnarray}
The linear independence of $\{M_i\}$ implies the linear independence of
the set of maps $\{M_i \rho_A M_j^\dagger\}_{i,j}$ on the density
matrices. For this reason, we have
\begin{equation}
\tr\,\rho_BL_j^\dagger L_i = \left\{
\begin{array}{ll}
\delta_{ij} & \textrm{for}~i,j \leq n \\
0 & \textrm{for}~i>n~\textrm{or}~j> n
\end{array}
\right.
\end{equation}
and consequently $\tr\,\rho_B L_i^\dagger L_i=0$ for all $i>n$. Since
$\rho_B$ is a canonical equilibrium state, it has full rank and therefore
$L_i=0$ for all $i>n$. This shows that the summation in
Eq.~(\ref{eq:U_AB_expansion}) contains only $n$ terms. The following
theorem directly follows from this expansion.
\begin{theorem}
\label{thm:gen1} Let $\mathcal{E}$ be the quantum operation given
in (\ref{eq:Kraus}) where $M_i$ are linearly independent.
\begin{itemize}
\item[(a)] If $Q$ is an HTO for $\mathcal{E}$, then there is an
$n\times n$ hermitian matrix $q$, (which might be called as the
\emph{heat transfer matrix}) such that
\begin{equation}
Q=\sum_{i,j=1}^n q_{ij} M_i^\dagger M_j \quad.
\label{eq:Q_q_relation}
\end{equation}
\item[(b)] If $n\geq2$ and the set of $n^2$ operators
$\{M_i^\dagger M_j\}_{i,j=1}^n$ is linearly independent, then $Q$
is an HTO for $\mathcal{E}$ if and only if the unique matrix $q$
in (\ref{eq:Q_q_relation}) satisfies
\begin{equation}
\tr\, e^{-q/k_B T} < 1~~~.
\label{eq:q_inequality}
\end{equation}
For $n=1$, $Q$ is an HTO if and only if $Q=\alpha\mathds{1}_A$
where $\alpha\geq0$.
\item[(c)] If $\{M_i^\dagger M_j\}_{i,j=1}^n$ is linearly
dependent, then the condition (\ref{eq:q_inequality}) is
sufficient for $Q$ given by (\ref{eq:Q_q_relation}) to be a
possible HTO; but that condition is not necessary.
\end{itemize}
\end{theorem}
Note that the set of possible quantum operations on A is a convex set. It
can be shown that the extreme points of this convex set correspond to
quantum operations for which the set of operators $\{M_i^\dagger
M_j\}_{i,j=1}^n$ is linearly independent\cite{BhatiaPos}. Therefore, part
(b) of the theorem above gives a complete characterization of the HTOs of
all extreme quantum operations.
\textit{Proof:}
For (a), the use of the definition of the HTO given in
Eq.~(\ref{eq:Q_definition}) directly leads to
Eq.~(\ref{eq:Q_q_relation}) where
\begin{equation}
q_{ij} = \tr \left(\rho_B L_i^\dagger H_B L_j\right)-E_B\delta_{ij}\quad.
\end{equation}
To show (b), first write the isometry condition $U_{AB}^\dagger
U_{AB}=\mathds{1}_{AB}$ as
\begin{equation}
\sum_{i,j=1}^n (M_i^\dagger M_j)_A \otimes (L_i^\dagger L_j)_B =\mathds{1}_A\otimes \mathds{1}_B~~.
\end{equation}
The linear independence of $\{M_i^\dagger M_j\}_{i,j=1}^n$ then leads to
the following relations
\begin{equation}
L_i^\dagger L_j = \delta_{ij} \mathds{1}_B\quad.
\end{equation}
In other words, each $L_i$ is an isometry on the bath that maps
$\mathcal{H}_B$ onto mutually perpendicular subspaces of $\mathcal{H}_B$.
The last set of relations is actually equivalent to the statement that
$U_{AB}$ is an isometry.
Consider now a hypothetical $n$-level system R. Let $\{\ket{i}\}_{i=1}^n$
be an orthonormal basis for the Hilbert space of R. The following operator
on the composite system RB,
\begin{equation}
W_{RB} = \sum_{i=1}^n (\ket{1}\bra{i})_R \otimes (L_i)_B
\label{eq:W_RB}
\end{equation}
is then an isometry. It can also be readily verified that $W_{RB}$
is an isometry if and only if $U_{AB}$ is also an isometry. The
most important point in here is that $W_{RB}$ is a realization of
a Landauer erasure on R. Furthermore, the HTO associated with this
erasure is given by
\begin{eqnarray}
Q^\prime_R &=& \tr_B \left[ \left(\mathds{1}_R\otimes \rho_B\right) W_{RB}^\dagger(\mathds{1}_R\otimes H_B )W_{RB} \right]
- E_B \mathds{1}_R\quad. \\
&=& \sum_{i,j} q_{ij} \ket{i}\bra{j}~~.
\label{eq:Qprime_R}
\end{eqnarray}
For $n\geq2$, the application of Theorem \ref{cor:Landauer} proves
the claim. The result for the special case of $n=1$ also follows
trivially from the discussion in this subsection.
Part (c) follows from the converse of the argument used above for
part (b). Suppose that $q$ satisfies (\ref{eq:q_inequality}). In
that case, $Q^\prime_R$ given by Eq.~(\ref{eq:Qprime_R}) is an
allowed HTO for a Landauer erasure. Hence one can construct the
associated realization; let $W_{RB}$ be the isometry for this
realization. We use Eq.~(\ref{eq:W_RB}) to define $L_i$ and then
use Eq.~(\ref{eq:U_AB_expansion}) to define $U_{AB}$. It is
straightforward to verify that $U_{AB}$ is the desired isometry
that realizes $\mathcal{E}$ and has $q$ as the heat transfer
matrix.
Finally note that any $q$ that satisfies (\ref{eq:q_inequality})
necessarily yields a positive definite $Q$. However, Theorem
\ref{thm:complete} allows HTOs that may have negative eigenvalues. This
shows that the condition (\ref{eq:q_inequality}) is not necessary. The
proof of Theorem \ref{thm:gen1} is thus completed. $\Box$
Note that part (b) of Theorem \ref{thm:gen1} is proved by finding
a closely related Landauer erasure on a hypothetical system. It is
possible to extend the scope of part (b) by making use of complete
erasures. However, the expressions used in such an extension do
not appear to be particularly illuminating and hence they are not
given in here. Theorem \ref{thm:gen1} is a good starting point for
discussing the properties of HTOs associated with generic quantum
operations. The following result is an example. It is essentially
a different way of saying that the set of HTOs associated with a
given quantum operation has no upper limit.
\begin{theorem}
Let $q$ be a heat transfer matrix for $\mathcal{E}$ which is
related to the corresponding HTO by (\ref{eq:Q_q_relation}). Then,
for any $q^\prime$ with $q^\prime>q$, $q^\prime$ is also a
possible heat transfer matrix.
\end{theorem}
To prove this, first define the matrix $s=q^\prime-q$ and observe that it
is strictly positive definite, i.e., all eigenvalues of $s$ are positive.
In such a case, it is possible to find a positive number $t\geq1$ such
that the matrix $q+ts$ satisfies (\ref{eq:q_inequality}) and therefore
$q+ts$ is a possible heat transfer matrix by part (c) of Theorem
\ref{thm:gen1}. Finally, by the convexity of the set of HTOs, the matrix
\begin{equation}
q^\prime = \frac{1}{t}(q+ts)+\frac{t-1}{t}q
\end{equation}
is a possible heat transfer matrix. This completes the proof.
$\Box$
Note that this theorem cannot be extended by replacing the
matrices $q$ by the associated HTOs mainly due to the restriction
in Theorem \ref{thm:gen1}~(a) which must be satisfied by all HTOs.
It is plausible that the strict inequality $q^\prime>q$ can be
replaced by the weaker requirement of $q^\prime\geq q$, but this
conjecture is still awaiting proof.
\section{Differences from Landauer's Bound}
\label{sec:differences}
The restrictions satisfied by the HTOs are closely related to LEP since
both follow only from the condition that the transformation on the
composite system is an isometry. However, LEP only relates the dumped heat
to the drop in the von Neumann entropy of the state of the device. For
complete erasures, it turned out that LEP also gives a complete
characterization of the HTOs. In other words, Theorems \ref{thm:complete}
and \ref{cor:Landauer} are identical in content with Theorem
\ref{thm:genlep}; they are just expressing the same statement using
different equations. As it will be shown below, this is not the case for
other quantum operations, i.e., it is possible to find examples of
operations $\mathcal{E}$ that are not complete erasures, such that the
HTOs associated with $\mathcal{E}$ satisfy further restrictions that
cannot be explained by LEP alone. Such operations can be considered to be
``partial erasures'' as some part of the initial information present in
the device will be retained (since $\mathcal{E}(\rho)$ depends on $\rho$).
The Landauer bound in Theorem \ref{thm:genlep} still applies to such
$\mathcal{E}$, but this bound fails to imply the additional features of
HTOs that are stated in Theorem \ref{thm:gen1}.
For example, the special structure of the HTOs given in Theorem
\ref{thm:gen1}~(a) does not appear in the generalized LEP inequality
(\ref{eq:entropic_LEP}). To see this clearly, consider unitary rotations
(or isometries) as quantum operations, i.e., $\mathcal{E}(\rho_A)=W\rho_A
W^\dagger$ where $W^\dagger W=\mathds{1}_A$. In such a case, the entropic
bound (\ref{eq:entropic_LEP}) merely states that the HTO must be a
positive-semidefinite operator. However, in reality, the HTOs should not
only be positive-semidefinite but they should also be a multiple of
identity as asserted in Theorem \ref{thm:gen1}~(b). Only in such a case,
the average heat dumped to the bath becomes independent of the state of
the device. Otherwise, any $Q$ which is not a multiple of identity would
create serious problems for quantum mechanics: it implies that by
investigating the energy of the bath, some information can be obtained
about the state of A without disturbing it! It is apparent that Theorem
\ref{thm:gen1}~(a) is intimately related with the characteristic features
of quantum information.
Apart from this, there are situations where a given $Q$ satisfies the
special form given in Theorem \ref{thm:gen1}~(a) and the entropic
inequalities (\ref{eq:entropic_LEP}), but it still fails to be an HTO. As
an example, consider the special problem of finding an HTO of the form
$Q=\alpha\mathds{1}_A$ in such a way that $\alpha$ has the minimum
possible value. If the quantum operation is extremal, then by Theorem
\ref{thm:gen1}~(b), $Q$ is an HTO if and only if $\alpha>k_BT\ln n$ where
$n$ is the minimum number of Kraus operators (for $n\geq2$). However, if
only the entropic inequalities (\ref{eq:entropic_LEP}) are used, it is
possible to obtain a different lower bound for $\alpha$, especially if the
quantum operation is near to identity.
As an example, consider the one-parameter family $\mathcal{E}_t$ of
quantum operations defined by $n=2$ Kraus operators
\begin{eqnarray}
M_1(t) &=& t X\quad,\\
M_2(t) &=& \sqrt{\mathds{1}-t^2 X^\dagger X}\quad,\\
\mathcal{E}_t(\rho_A) &=& \sum_{i=1}^2 M_i(t)^\dagger \rho_A M_i(t) \quad,
\end{eqnarray}
where $X$ is a non-hermitian operator such that $\{M_i^\dagger(t)
M_j(t)\}_{i,j=1}^2$ are linearly independent. Note that in the limit
$t\longrightarrow0$, the operation $\mathcal{E}_t$ approaches to the
identity transformation. Therefore, the maximum drop in the von Neumann
entropy of the device,
\begin{equation}
B_t = \max_{\rho_A} \left( S(\rho_A)-S(\mathcal{E}_t(\rho_A)) \right)\quad,
\end{equation}
approaches zero in the same limit. Then, the smallest possible value of
the coefficient $\alpha$ allowed by the entropic restrictions of LEP is
$\alpha_{\textrm{min}}=k_BT B_t$, which also tends to zero as
$t\longrightarrow0$. However, it is shown above that $\alpha>k_BT\ln2$,
independent of how close $\mathcal{E}_t$ to the identity operation. This
clearly shows that Theorem \ref{thm:gen1}~(b) contains restrictions on
HTOs that are not implied by the entropic LEP restrictions.
In order to eliminate a possible misunderstanding that may arise from the
previous example, let us stress on the fact that once we lift the
restriction that $Q$ is proportional to the identity, it is possible to
find HTOs that approach to the zero operator when $\mathcal{E}_t$
approaches to the identity operation. For example, for $\mathcal{E}_t$
given above,
\begin{equation}
Q=\sum_{i=1}^2 q_i M_i(t)^\dagger M_i(t)\quad,
\end{equation}
with $\beta q_2=t^2$ and $\beta q_1\approx \ln(1/t^2)$ yields an HTO $Q$
with
\begin{equation}
\beta Q\approx t^2 \left(\mathds{1}_A +\ln(1/t^2) X^\dagger X\right)
\end{equation}
which converges to zero as $t\longrightarrow0$.
\section{Conclusions}
\label{sec:conclusions}
In this article, the concept of heat transfer operator is
introduced and some of its properties are discussed. The HTO is
the main quantity needed for analyzing the energetics of quantum
operations. Hence, the complete characterization of the HTOs
associated with a given operation $\mathcal{E}$ occupies a central
place in the thermodynamics of quantum information processing.
This article mainly contains results on this problem. A complete
characterization could be provided only for complete erasures and
extreme operations. The characterization of HTOs of generic
non-extremal quantum operations is still an open problem.
The set of HTOs associated with a given quantum operation might be
considered as a means of describing the disturbance caused by the
operation on the environment. This description is provided through a set
of operators that has a clear operational interpretation. This points out
the main theoretical interest for this problem. In addition to this, both
the results obtained and the techniques used in this article can be useful
in the implementation of non-unitary operations in future quantum
computers. For example, the minimization of the average heat emission
during an erasure operation requires a careful construction of the
transformation on the extended system of the device and the bath. The
constructions used in this article will be useful for such design
problems.
\section*{Acknowledgements}
The authors are grateful to N. K. Pak for stimulating discussions.
|
\section{Introduction}
Of all physical theories about Nature, only two branches are
independent of concrete details of the systems being considered, i.e.
thermodynamics and relativity (in either the special or general sense). These
two theories are about the universal principles every
physical system must obey and are hence referred to as principle
theories. All other theories are about concrete systems. It is puzzling why in
the first place Nature prefers to abide by {\em two principle theories} rather
than simply one (puzzle of two), because two is more dangerous in the
sense that potential contradictions between them might occur. However,
Nature is quite smart in avoiding such contradictions. Even more, in many
cases the two principle theories seem to be mysteriously connected
to each other. This connection is especially transparent when one talks
about relativity in the general sense: on the one hand, for every black hole
solution of the general theory of relativity one can define thermodynamic
quantities like entropy and temperature and check that the classical laws of
thermodynamics are obeyed; on the other hand, it has repeatedly been
discussed that the Einstein equation \cite{Jacobson:1995p4065}
\cite{Padmanabhan:2009p5086} \cite{Padmanabhan:2009p5122} and the
action \cite{Caravelli:2010p5148} of the general theory of relativity can be
derived from the first law of thermodynamics, and recently Verlinde
\cite{Verlinde:2010p5107} went
further by proposing that gravity is an entropic force and thus calling for an
end of gravity as a fundamental force. Subsequent surge of works appeared
almost instantly, including \cite{Smolin:2010p5123}, which applied
Verlinde's idea in loop quantum gravity,
\cite{Padmanabhan:2010p5128} and \cite{Cai:2010p5127}, both derived
the Friedman equation using Verlinde's idea,
\cite{Wang:2010p5132}, which applied Verlinde's idea in establishing
UV/IR relation and obtaining holographic dark energy,
\cite{Wang:2010p5136}, which tries to interpret electrostatic force also as an
entropic force, \cite{Zhang:2010p5140}, \cite{Ling:2010p5141}
and \cite{Wei:2010p5145}, which applied Verlinde's ideas in
the settings of modified gravity, brane cosmology and Horava gravity
respectively, and \cite{Wang:2010p5144}, which proposed two novel
kind of cosmic perturbations corresponding to Verlinde's emergent gravity.
The idea of emergent gravity also attracted some quantum
information people, see \cite{Lee:2010p5153} for an alternative proposal.
Verlinde's proposal is certainly very beautiful. If proven true, a significant
portion of modern theories of fundamental physics should be reformulated.
It not only has the potential of addressing the long standing problem of
quantum gravity by identifying gravity as a purely macroscopic effect, but
also implies that space itself must be emergent as a macroscopic effect.
Moreover, it has the potential to resolve the puzzle of two mentioned
above. However, Verlinde's idea also raises many new problems because
most concepts taking space as a fundamental existence must be
reexamined. Verlinde himself has
noticed this point and explained in particular the concepts of inertia and
Newton's law of classical mechanics from the new point of view.
In this article I will comment that there are more problems which should
be checked against before one can take Verlinde's proposal more seriously.
The origin of the principle of relativity is
among these problems. It is found that there is a large group of hidden
symmetries for thermodynamics which has not been discussed before. By
incorporating this hidden symmetry group together with Verlinde's idea and
the second law of thermodynamics, it is shown that the principle of
relativity
arises naturally. Meanwhile, a clearer understanding about the role of
holographic principle in Verlinde's proposal is given. It turns out that
holography is a requirement of democracy between different scaling
dimensions and the fact that space is three dimensional.
\section{Verlinde's proposal: problems to be answered}
Verlinde's proposal contains the following key ingredients:
\begin{itemize}
\item space is emergent, the part of space which has not yet emerged
is enclosed by a holographic screen and the entropy is proportional to the
area of the screen;
\item gravity is an entropic force just like any other generalized forces
entering in the first law of thermodynamics. More concretely, gravity
is caused by the change of entropy behind the holographic screen due to
the emergence of space;
\item the temperature is either related to the acceleration of the observer via
Unruh's law or related to the total energy of the system via equipartition
law (the equipartition of energy on black hole horizons is discussed
earlier by Padmanabhan in \cite{Padmanabhan:2003p5146}, see also
\cite{Padmanabhan:2009p5086}),
and the total energy of the system is also equal to the mass behind the
holographic screen times $c^2$, i.e. $E=Mc^2$. Using Unruh's law for
temperature leads to Newton's law of classical mechanics, and using
equipartition law and $E=Mc^2$ gives rise to Newton's law of gravitational
force.
\end{itemize}
A lot more have been discussed in Verlinde's paper
\cite{Verlinde:2010p5107}, however besides
the above new ingredients the rest arguments and sketchy derivations seem to
have been discussed in or implied by earlier works since Jacobson's
\cite{Jacobson:1995p4065}, in
particular, the formal derivation of Einstein equation is now new.
Now let us take a more careful look at the new ideas listed above.
First of all, given that space is emergent, one has to explain why is it look
like what we perceive it. In particular, why is space three dimensional, or
even why the dimension of space is an integer? This is a problem one has to
answer in an emergent theory of space but these were not paid for a single
word in Verlinde's paper;
Second, it is known since the 1930's \cite{Tolman:1934} that classical
thermodynamics is
incompatible with the special theory of relativity if spacetime is considered as
a fundamental existence. However, in an emergent theory of space one has to
see the principle of relativity also as an emergent consequence. It remains to
check whether the emergence of the principle of relativity is possible or not
following Verlinde's proposal, and I think this is by now one of the most
severe obstacle to be overcome before this new proposal can be more widely
accepted in the physics community;
Third, in deriving the Newton's law of gravity, the equipartition law together
with the relation $E=Mc^2$ were used as input. But the origin of the latter
relation is only known from special relativity. It gives an impression that
special relativity is provided as another fundamental assumption, but even
though it is still strange that in order to obtain the nonrelativistic law for
gravity one has to use the relativistic relation for the total energy. Also the
Unruh temperature relation is a consequence of relativistic motion, while it is
used in deriving the nonrelativistic law of classical mechanics. Moreover,
the use of equipartition law also feels strange, because equipartition law
depends on the microscopic details of the system while Newton's law doesn't.
There are other issues which need more interpretations but I prefer not to
concentrate on them and just emphasis on the problem of emergence of the
principle of relativity. If the principle of relativity is put in by hand, it will
seriously hurt the picture of the emergence of space, and the puzzle of two
will not be resolved. On the contrary, if the
principle of relativity is emergent, then not only the puzzle of two is
resolved, it will also call for an improvement for the derivation of the
Newton's law of gravity, or simply view Newton's law as the nonrelativistic
limit of general relativity as usual while considering general relativity as an
emergent theory about gravity. In either way the puzzle of deriving
nonrelativistic laws using relativistic formulas without taking a
nonrelativistic limit should be avoided.
In the next section I will propose a possible solution to the emergence of
the principle of relativity by unraveling a large group of hidden symmetries
of classical thermodynamics. Assuming that space is emergent and plays as extensive variables in the first law of thermodynamics, it will be seen that the Poincare group is naturally contained in the hidden symmetry group of thermodynamics, which gives an explanation of the principle of relativity as an emergent concept from thermodynamic geometry.
\section{Geometry and hidden symmetries of classical
thermodynamics}
There are quite a few different prescriptions of thermodynamics as a
geometric system. Gibbs \cite{J.Gibbs} described the thermodynamic
phase space as a contact manifold. Weinhold \cite{F. Weinhold} and
Ruppeiner \cite{G. Ruppeiner} respectively described the geometry of the
thermodynamic configuration space (i.e. the space of equilibrium states) as
Riemannian geometries (with different metric choices). And recently H.
Quevedo \cite{Quevedo:2006p1979} \cite{Quevedo:2007p1262} described
the geometry of the thermodynamic phase space as a
contact Riemannian geometry invariant under the group of Legendre
transformations.
For my purpose it is tempting to describe the geometry of the
thermodynamic phase space as a contact, Riemannian geometry whose pull
back to the thermodynamic configuration space has a metric with
Lorentzian signature.
To begin with, let me describe the contact geometry of the thermodynamic
phase space $\mathcal{T}$, following the spirits of Gibbs \cite{J.Gibbs}.
The space $\mathcal{T}$ is a $2n+1$ dimensional contact
manifold with coordinates $(\Phi, E^a, I_a)$ ($a=1, 2, ..., n$), where
$\Phi$ is a thermodynamic potential, $E^a$ and $I_a$ are respectively
extensive and intensive variables of the system, and a contact one form
$\Theta = d\Phi - \sum_a I_a dE^a$ obeying $(d\Theta)^{\wedge n}
\wedge \Theta \neq 0$ is needed in order to identify the contact structure
of the manifold $\mathcal{T}$. The space
of classical equilibrium thermodynamical states $\mathcal{E} =
\mathrm{span \ of \ }\{E^a\}$
is embedded in $\mathcal{T}$ as a subspace by a smooth mapping
\[
\varphi: \mathcal{E} \rightarrow \mathcal{T}, \qquad
\varphi (E^a) = (\Phi(E_a), E^a, I_a(E^a)),
\]
and the pull back condition $\varphi^\ast(\Theta) =0$ gives rise to the first
law of thermodynamics, i.e.
\begin{equation}
d\Phi =\sum_a I_a dE^a, \label{1st}
\end{equation}
together with the equations of states
\begin{equation}
I_a = \frac{\partial \Phi}{\partial E^a}. \label{2nd}
\end{equation}
It is known from standard texts on differential geometry that a contact
manifold possesses a very special group of symmetries, i.e. the group of
contact transformations. The group of Legendre transformations, i.e.
\begin{align*}
\Phi = \tilde{\Phi} - \sum_{a\in \mathcal{I}'} \tilde{I}_a \tilde{E}^a,
\quad
E^a = - \tilde{I}_a, \quad
I_a = \tilde{E}^a,
\end{align*}
for $a \in \mathcal{I}'$ with $\mathcal{I}=\{0,1,2,...,n-1\}$ and
$\mathcal{I}' \subset \mathcal{I}$ is any subset therein, is a subset of
the group of contact transformations, and H.Quevedo stressed very much
on the invariance under this subgroup in his prescription of
geometrothermodynamics \cite{Quevedo:2006p1979}
\cite{Quevedo:2007p1262}. It is however not my intention to keep my
eyes addicted to the group of Legendre transformations. Rather, the
introduction of a Riemannian metric on $\mathcal{T}$ seems more
attractive. The concrete form of the metric on $\mathcal{T}$ is not important. What is important is that the pull back of this metric naturally introduces a metric on the thermodynamic configuration space $\mathcal{E}$, which, in its most general form, can be written as
\begin{equation}
ds^{2} = g_{ab}(E) dE^{a} dE^{b}. \label{dsmetric}
\end{equation}
Now one of the crucial part of this work turns up. There exists a large
continuous group of hidden symmetries which keeps both
(\ref{1st}) and (\ref{dsmetric}) invariant. This hidden symmetry group
is the group ${G}$ of general coordinate transformations in the space $
\mathcal{E}$, i.e.
\begin{align*}
E^a &\rightarrow E^{\prime a} = E^{\prime a}(E^b), \quad \Phi
\rightarrow \Phi'=\Phi,\\
dE^a &\rightarrow \left(\frac{\partial E^a}{\partial E^{\prime b}}\right)
dE^{\prime b},\qquad
I_a \rightarrow \left(\frac{\partial E^{\prime b}}{\partial E^{a}}\right)
I'_{ b}.
\end{align*}
In the above, the transformation law of $I_a$ is triggered by the relation
(\ref{2nd}).
Borrowing some terminologies from the general theory of relativity, $dE^a$
transform under the group ${G}$ as a contravariant vector, $I_a$ transform
as a covariant vector, while $\Phi$ transforms as a scalar.
Notice that although in the above I am using geometric terminologies like
manifold, Riemannian metric, contravariant and covariant vectors etc,
no fundamental existence of space is assumed actually. As Verlinde has
emphasized, the definition of thermodynamics does not need the existence
of space. Space actually emerges as macroscopic consequences as some of
the extensive variables (generalized displacements) in the space
$\mathcal{E}$. Therefore, the group $G$ is by now purely
thermodynamic in nature, nothing
relevant to spacetime symmetry has entered into play.
The invariance under the group $G$ is totally a new
observation which has not been discussed before. Considering the fact that
there has already been quite some works \cite{F. Weinhold}
\cite{G. Ruppeiner} assigning a Riemannian metric to
the space of equilibrium states of thermodynamics, it is quite strange why
this large group of isometries have not been discussed before in the
context of thermodynamic geometries (for the case of Weinhold geometry,
the isometry group is discussed in \cite{P. Salamon}, but the groups
considered there is not the same as the one discussed here. Meanwhile, the
deep implications implied by the symmetries is somehow not widely
acknowledged). Anyway, the finding of this large group of
hidden symmetries provides room for an
interpretation of the principle of relativity as an emergent effect from
thermodynamics, as will be seen below.
First let me fix the choice of the thermodynamic potential $\Phi$ by
identifying it as the total internal energy of the system. This breaks the Legendre
symmetry but leaves the $G$ invariance unaffected. Meanwhile, this
specific choice of thermodynamic potential implies that the entropy $S$ is
among the extensive variables, let me give it the index $0$, i.e. $E^{0}=S$.
Unlike the case of Weinhold and Ruppeiner, I did not associate the metric
$g_{ab}(E)$ with the hessian of any specific thermodynamic function, so
there is still enough freedom in choosing the signature of $g_{ab}(E)$. Let
the signature be Lorentzian and let $E^{0}$ be the coordinate bearing the
different signature from others. The $G$ symmetry allows to make different
choices for the reference frames on $\mathcal{E}$. Near any point $X$ in
$\mathcal{E}$, take a small neighborhood $U$ thereof, then using
$G$-transformations one can always fix the metric (\ref{dsmetric}) on $U$
to a special form, i.e.
\begin{equation}
ds^{2} = \eta_{ab} dE^{a} dE^{b}, \label{dsinertial}
\end{equation}
where $\eta_{ab}$ is the Lorentzian metric in $n$ dimensions.
This is the analogue of taking a local inertial frame in the general theory of
relativity, but now practiced purely in the framework of thermodynamics.
Once the frame (\ref{dsinertial}) is taken, the $G$ symmetry is broken, but a residual subgroup $H$ survives, which consists of linear
transformations among $E^a$, i.e.
\begin{equation}
E^a \rightarrow E^{\prime a} =\Lambda^a{}_b E^b + T^a,
\label{linear}
\end{equation}
in which $\Lambda^a{}_b$ belongs to the orthogonal group $SO(n-1,1)$
and $T^a$ are constants. Clearly eq.(\ref{1st}) is
invariant under such a group of transformations, provided $I_a$ transform
inversely under $H$.
Readers may have already noticed the close analogue of eq.(\ref{linear})
with the Poincare group. To actually establish the relationship between the
group $H$ and the Poincare group, extra input is needed, including the
second law of thermodynamics and Verlinde's proposal for an emergent
space.
Consider first the second law of thermodynamics. There are many
presentations for the second law in the literature. For convenience, we take
the following presentation:
Second law: {\em In any thermodynamic process connecting two
equilibrium states of an isolated macroscopic system, the entropy does not
decrease, \i.e.}
\[
dS \geq 0.
\]
This presentation can also be put in another way, i.e. {\em any equilibrium
state of a given macroscopic system of lower entropy cannot be the
consequence of a thermodynamic process of another equilibrium state of
higher entropy.} Such a statement is reminiscent to the causality principle
of relativistic physics and may be called the thermal causality principle.
Careful readers may have felt uneasy with the last paragraph.
According to eq.(\ref{linear}), $S$ is not a scalar under the action of the
group $H$. In other words, there is not a unique choice for the extensive
variable $S$ in the presence of the symmetry group $H$. So the thermal
causality principle must
be formulated in an invariant way under the action of the group $H$. For
this purpose we need an invariant quantity under the action of $H$, and
the line element on the space $\mathcal{E}$ happens to fill this
gap.
Geometrically the line element (\ref{dsinertial}) describes the invariant
distance between two points in $U$. What is the thermodynamic meaning
of such a distance? It is clear that the two points connected by $ds^{2}$
correspond to two distinct equilibrium states, so the line element between
them must corresponds to a thermodynamic process evolving from one
state to the other. Now the following crucial question arises: given any two
states $A, B$ in $U$, does the thermodynamic process connecting them
always exist? To answer this question, let me first fix a reference frame on
$U$ such that $S_{A} \leq S_{B}$. Then the second
law implies that there can possibly be a thermodynamic process
evolving from the state $A$ to the state $B$. However, such a process is
not guaranteed to exist, since a change of frame on $U$ can
spoil the inequality $S_{A} \leq S_{B}$. So, if $A$ and $B$ are such that
$S_{A}< S_{B}$ in one frame but $S_{A}>S_{B}$ in some other frame,
the presumed thermodynamic process evolving from $A$ to $B$ (or vise
versa) should be excluded by the second law. In other words, such states
$A$ and $B$ must not be causally connected via thermodynamic process.
On the other hand, it is possible that for properly chosen states $A$ and $B
$, the inequality $S_{A} \leq S_{B}$ holds for all allowed choices of
frames on $U$. In such cases, one cannot exclude the
possibility that $B$ is the thermodynamic consequence of $A$ through
some thermal process. It is not a hard practice to show that $S_{A}
\leq S_{B}$ holds for all allowed choices of frames on $U$ if and only if
$S_{A} \leq S_{B}$ holds in one frame on $U$ and the line element $ds_
{AB}^{2}$ between $A, B$ obeys $ds_{AB}^{2}\geq 0$. So,
the thermal causality principle can be formulated as follows:
Invariant presentation of the second law: {\em In any thermodynamic process connecting two equilibrium states of
an isolated system, the line element $ds^2$ must be nonnegative.}
For this reason we can possibly call $ds$ the {\em proper entropy change}
between the two states. Notice that I did not assign an arrow to the line
element, so a thermodynamic process corresponds to both $ds_{AB}
\geq0$ and $ds_{AB} \leq0$.
Now let me follow Verlinde's idea and consider the whole universe as an
emergent macroscopic system. This implies, among other things, that the
spacial coordinates $X^i$ are among the
extensive variables $E^a$ in the space $\mathcal{E}$. Moreover, since
nothing is assumed to exist outside the
universe, one can consider the universe as an isolated system, i.e. the
condition for the thermal causality principle hold, any physical
thermodynamic process must obey $ds^{2} \geq 0$.
Unlike Verlinde's original proposal for a single emergent dimension, I
postulate here that {all the spacial dimensions are emergent}. With the
emergent spacial coordinates as extensive variables, the first law (\ref{1st})
should be modified as follows:
\begin{equation}
dE = TdS - F_i dX^i - pdV + \mu dN, \label{law}
\end{equation}
where the sum over $i$ extends through all spacial dimensions. Note that
$F_i$ are emergent forces just like $p$ does, and $X^i$ and $V$ are all
macroscopic quantities which need not have an microscopic origin. This is
what the term emergence of space means.
Since the spacial coordinates are identified as
generalized displacements, the total number of such displacements must be
an integer. This explains why
the dimension of space is an integer. It is natural to assign a scaling
dimension for each of the extensive variables. Doing so one sees that only
when space is three dimensional and $S$ is proportional to the area of a
holographic screen, the right hand side of eq.(\ref{law}) can be
{\em democratic} (i.e. evenly distributed) between quantities of different
scaling dimensions. In the above, $N$, the number of total microscopic
degrees of freedom, is zero dimensional, $X^i$
are one dimensional, $S$ is two dimensional and $V$ is three dimensional.
In this way the role of holographic principle in Verlinde's proposal and the
implicit assumption of $3$ spacial dimensions are replaced by a single
requirement of democracy between scaling dimensions.
In a process with $dV=dN=0$, the action of the subgroup
$P$ of $H$ leaving $V$ and $N$ invariant is isomorphic to the
Poincare group in the subspace $\mathcal{M}$ of $\mathcal{E}$ spanned
by $S$ and $X^i$. If I assume that $dS$ is proportional to the time
elapsed between the two equilibrium states (and this has to be so, because
in any reference frame, $S$ is a monotonic function of time $t$, thus $dS
\propto dt$ locally, and the procedure of fixing the metric (\ref{dsinertial})
could fix the constant of proportionality to 1), then the relativistic causality
principle naturally follow from the thermal causality
principle.
What is the role of $H$ in the process involving non zero $dV$ or $dN$?
The answer could also be very interesting. For instance, if a process
involves non zero $dV$, then the group $H$ is bigger than $P$, i.e. the
pull back line element $ds^2$ possesses a bigger isometry group. Such a
group provides room for cosmological expansion, and the corresponding
intensive variable $p$ could possibly play the role of dark energy. However,
at present, such possibilities must be regarded as speculative, because much
more work on the understanding of thermodynamic emergence of space
has to be done before such speculations can be made more sounded. As
for $dN \neq 0$, the process will involve production or destruction of
particles and that is beyond the present discussion on the emergence of the
principle of relativity.
\section{Discussions}
Using the hidden symmetry group of thermodynamics and the thermal
causality principle it is found that the principle of relativity arises
naturally as an emergent consequence of thermodynamics. Thermal
causality is identified with the relativistic causality by requiring
that the increase in entropy, $dS$, is proportional to the time elapse. This
analysis resolves a number of problems left over from Verlinde's proposal
for an emergent space and gravity.
One may wonder why in the first place the principle of relativity could arise
as an emergent consequence of thermodynamics which has long been
known to be incompatible with special relativity regarded as a fundamental
principle theory. In particular, why is the total energy $E$ taken as the
thermodynamic potential transform as a scalar under the emergent Poincare
group, while in traditional special relativity, energy transforms as the zeroth
component $p^0$ of the energy-momentum 4-vector. The answer is that
$p^0$ and $E$ are completely different objects which should be not
confused with each
other. Actually, $p^0$ refers to the energy of certain microscopic degree of
freedom, while $E$ counts the total energy of all microscopic degrees of
freedom in the system, so by definition $E$ is an integral quantity in
which the covariant and contravariant actions of Poincare group cancel
completely. Due to the same reason, there is no reason to write $E=Mc^2$,
because $Mc^2$ only represents the energy of a microscopic
degree of freedom of mass $M$ at rest, while $E$ counts the energy of all
microscopic degrees of freedom, and most of these are not at rest.
Another reason supporting the view point of not regarding $E$ as a
component of relativistic $4$-vector comes from the new relationships
between relativity and thermodynamics. If spacetime were fundamental and
$E$ is the total energy of a thermodynamic system moving in space, then it
seems inevitable that $E$ should change under relativistic change of
coordinates. However, in the present picture, {\em space itself is emergent
from thermodynamics}. The thermodynamic system no longer moves in
space, and $E$ contains not only the energy of ordinary matter but also the
energy of space. (The picture of space as emergent from
thermodynamics implies that space itself can be heated, and thus stores
energy. On this point, I am grateful to T.~Padmanabhan for bringing my
attention to the paper \cite{Padmanabhan:2010} after the first versions of
the present paper have appeared on arXiv.) We human beings have no
previous
knowledge at all on how the total energy of matter and space changes
under coordinate changes in space and time.
In this article, only the linear subgroup of the hidden symmetries of
thermodynamics is analyzed in detail. A more thorough analysis on the
complete hidden symmetry group of thermodynamics should be considered
later, which is expected to reveal how general relativity arises from
thermodynamics. Whatever the details will be, I expect that a better
derivation of Einstein equation should follow and Newton's law of gravity
should arise as nonrelativistic limit of general relativity as usual, thus
removing the puzzle of deriving nonrelativistic laws of gravity from relativistic
formulas.
The argument made in this article is very preliminary. Much more works are
yet to be done. Among other things the hidden symmetry group $G$ of
thermodynamics is bigger than the general coordinate transformations for
general relativity provided in the first law of thermodynamics $dV$ and $dN$
are not simultaneously zero. The role of the extra symmetries is yet to be
understood, and perhaps this will give novel insights into the general theory
of relativity.
\section*{Acknowledgment}
This work is supported by the National Natural Science Foundation of
China (NSFC) through grant No.10875059.
\providecommand{\href}[2]{#2
|
\section{Introduction}
\label{sec:intro}
Transactional events (TE) provide powerful message-passing facilities
for concurrent programming \cite{te-hs}. TE extends Concurrent ML
(CML) \cite{cml-book} with a sequencing primitive \ttt{thenEvt}, which
allows an arbitrary number of sends or receives per event.
\ttt{thenEvt} is powerful enough that programmers can implement patterns
such as $n$-way rendezvous (synchronizing multiple threads at once) or
guarded receive (successfully synchronizing only if a received value
satisfies a predicate).
In previous work \cite{te-hs, te-ml}, we and other researchers
explored the semantics and implementation of transactional events.
However, no existing work discusses practical programming idioms for
TE. We have found in the course of our research that some each common CML
idioms are surprisingly difficult to reproduce in TE\@. This paper
presents our solutions to these problematic idioms, which include
client-server protocols and programs using CML's \ttt{wrap},
\ttt{guard}, and \ttt{wrapAbort}.
TE is implemented for Haskell and Caml. In our examples, we use Caml,
which is call-by-value.
\begin{SaveVerbatim}{verb:types}
type 'a chan
type 'a event
val newChan : unit -> 'a chan
val sync : 'a event -> 'a
val sendEvt : 'a chan -> 'a -> unit event
val recvEvt : 'a chan -> 'a event
val chooseEvt : 'a event -> 'a event
-> 'a event
val thenEvt : 'a event -> ('a -> 'b event)
-> 'b event
\end{SaveVerbatim}
\begin{SaveVerbatim}{verb:types2}
val alwaysEvt : 'a -> 'a event
val neverEvt : 'a event
val guard : (unit -> 'a event)
-> 'a event
val wrap : 'a event -> ('a -> 'b)
-> 'b event
val wrapAbort : 'a event
-> (unit -> unit)
-> 'a event
\end{SaveVerbatim}
\section{Background}
\label{sec:bg}
\begin{figure}[t]
\fbox{
\begin{minipage}{.54\textwidth}
{\BUseVerbatim{verb:types}}
\end{minipage}
\begin{minipage}{.42\textwidth}
{\BUseVerbatim{verb:types2}}
\end{minipage}
}
\caption{Types for several key CML/TE functions.}
\label{fig:types}
\end{figure}
TE \cite{te-hs,te-ml} and CML \cite{cml-book} are closely related
paradigms for concurrent programming that have been implemented for
several languages. This section briefly reviews TE and explains how
it differs from CML\@.
In CML, threads send values on channels. A channel of type
\ttt{'a chan} carries values of type \ttt{'a}. Communication is
\emph{synchronous}: a send blocks until matched with a receive in
another thread.
An {\em event} is a description of a communication to be
performed. The functions \ttt{sendEvt} and \ttt{recvEvt} produce
events that describe sends and receives, respectively.
\emph{Synchronizing} on an event with the function \ttt{sync} performs
the event; \ttt{sync} has type \ttt{'a event -> 'a}. Events may be
composed of other events. For example, the function \ttt{chooseEvt}
constructs an event that, when synchronized on, performs exactly one
of two events. The types of these and other functions appear in
Figure \ref{fig:types}.
TE extends CML with a new function \ttt{thenEvt}, which allows the
sequencing of events. Synchronizing on \ttt{thenEvt ev f} does the
following: (1) synchronize on \ttt{ev} to produce a result \ttt{v};
(2) apply \ttt{f} to \ttt{v} to produce a new event \ttt{ev2}; (3)
synchronize on \ttt{ev2} to produce a final result. Critically, these
three steps are all-or-nothing: if the second event cannot
successfully synchronize, then the first event does not (appear to)
happen. Therefore, we say that events built using \ttt{thenEvt} are
\emph{transactional}. A single event may communicate with multiple
threads, so the success of one synchronization may entail the success
of an arbitrary number of synchronizations in other threads.
Two more CML/TE functions are useful in combination with
\ttt{chooseEvt} and \ttt{thenEvt}. \ttt{alwaysEvt} is the event that
always succeeds: \ttt{sync (alwaysEvt x)} is equivalent to \ttt{x}.
\ttt{neverEvt} is the event that never succeeds: \ttt{sync neverEvt}
blocks forever.
The following example tries to do either a single send, or a receive
followed by a send:%
\begin{verbatim}
sync (chooseEvt (sendEvt c1 5)
(thenEvt (recvEvt c2) (fun x -> sendEvt c3 x)))
\end{verbatim}
\section{Server loops}
\label{sec:server}
A common concurrent programming idiom is a server thread that repeatedly handles requests from multiple clients. In CML, a server is often implemented as an infinite loop that calls \ttt{sync} at every iteration. In this section, we discuss why TE requires more sophisticated servers and how to implement them.
Consider the following function, which spawns a new thread to act as a server. The server sends increasing integers on a channel so that clients get a unique integer every time they receive on the channel.
\begin{verbatim}
let simpleIncrementServer () =
let c = newChan () in
let rec f y = sync (sendEvt c y); f (y + 1) in
Thread.create f 0; c
\end{verbatim}
A simple client receives on the server's channel.
\begin{verbatim}
let simpleIncrementClient c = sync (recvEvt c)
\end{verbatim}
Both client and server are perfectly valid as both TE and CML. However, in TE \ttt{thenEvt} can be used to write other clients that interact with this server in unexpected ways. The following code receives two integers and adds them together in a single event:
\begin{verbatim}
let complexIncrementClient c =
sync (thenEvt (recvEvt c) (fun x ->
thenEvt (recvEvt c) (fun y ->
alwaysEvt (x + y))))
\end{verbatim}
If this client and the server were to synchronize on their events, neither would succeed. The client could receive one integer from the server. The server event would block waiting for the client event to complete, as they would be participating in the same transaction. Meanwhile, the client would block waiting for another integer. Both sides would be stuck, so this particular client cannot successfully synchronize with the simple increment server.
A similar problem arises when {\em two} client threads receive an integer from the server and communicate with each other in the same event. A transaction consisting only of these two events and the server event cannot succeed. The server would need to send to both threads, but the server's event does one send.
To solve this problem, we need a server that can perform multiple sends in a single call to \ttt{sync}. In the TE code below, the server synchronizes on one event that can do an arbitrary number of sends. After each send, the event chooses between returning the sent value or recursively calling the server function.
\begin{verbatim}
let betterIncrementServer () =
let c = newChan () in
let rec evtLoop x =
thenEvt (sendEvt c x)
(fun _ -> chooseEvt (alwaysEvt (x + 1)) (evtLoop (x + 1))) in
let rec serverLoop x = serverLoop (sync (evtLoop x)) in
Thread.create serverLoop 0; c
\end{verbatim}
However, this problem can occur with many different server and client combinations. A better solution would be to create a generic combinator for an event that is repeated one or more times.
For this purpose, we define a function, \ttt{serverLoop}, suitable for
creating servers. It takes two arguments and loops forever. The
first argument is a pair, \ttt{(ev, b)}, of an event and any value.
The second argument is a function, \ttt{f}. After synchronizing on
\ttt{ev} to produce a value \ttt{a}, \ttt{serverLoop} calls \ttt{f} on
the pair \ttt{(a,b)} to produce a new \ttt{(ev,b)} pair, with which it
recurs. \ttt{b} acts as a loop-carried state for
\ttt{serverLoop}. Programmers can use \ttt{serverLoop} much like they
use \ttt{fold} to process lists.
Internally, \ttt{serverLoop} uses a second function, \ttt{evtLoop}, that creates an event that synchronizes on \ttt{ev} and then nondeterministically chooses between returning or recurring with the result of calling \ttt{f} to produce a new event and loop-carried state. In other words, \ttt{evtLoop} does exactly what \ttt{serverLoop} does but, crucially, in a single transaction.
Overall, \ttt{serverLoop} sequentially synchronizes on the events
computed by \ttt{f}, but transactions may end (starting the next transaction)
at any point in the sequence. Thus, clients can communicate with the
server any number of times in one synchronization.
\begin{verbatim}
(* evtLoop :
('a event * 'b) -> (('a * 'b) -> ('a event * 'b)) -> ('a * 'b) event *)
let rec evtLoop (ev, b) f =
thenEvt ev (fun a -> chooseEvt (alwaysEvt (a, b)) (evtLoop (f (a, b)) f))
(* serverLoop : ('a event * 'b) -> (('a * 'b) -> ('a event * 'b)) -> 'c *)
let rec serverLoop (ev, b) f = serverLoop (f (sync (evtLoop (ev, b) f))) f
\end{verbatim}
Using \ttt{serverLoop} to construct an increment server is straightforward.
We spawn a new thread to run \ttt{serverLoop} called with (1)
an initial event paired with the initial loop-carried counter and (2)
a function to construct the next event and counter by incrementing the
counter and creating the next send event.
\begin{verbatim}
let incrementServer () =
let c = newChan () in
Thread.create (serverLoop ((sendEvt c 0), 0))
(fun (_, x) -> (sendEvt c (x+1), x+1)); c
\end{verbatim}
\section{\texttt{wrap} and \texttt{guard}}
\label{sec:wrap}
\ttt{wrap} and \ttt{guard} (see Figure \ref{fig:types}) add post- and
pre-processing, respectively, to CML events. \ttt{sync (wrap ev f)}
synchronizes on \ttt{ev}, then applies \ttt{f} to the result.
\ttt{sync (guard g)} synchronizes on the result of \ttt{g ()}. In
this section, we discuss how to adapt programs that use these
functions to TE.
The following code uses \ttt{wrap} to perform two receives in either
order:
\begin{verbatim}
sync (chooseEvt (wrap (recvEvt c1) (fun x -> (x, sync (recvEvt c2))))
(wrap (recvEvt c2) (fun x -> (sync (recvEvt c1), x))))
\end{verbatim}
Suppose we were to rewrite this code using \ttt{thenEvt}:
\begin{verbatim}
sync (chooseEvt
(thenEvt (recvEvt c1)
(fun x -> thenEvt (recvEvt c2) (fun y -> alwaysEvt (x,y))))
(thenEvt (recvEvt c2)
(fun y -> thenEvt (recvEvt c1) (fun x -> alwaysEvt (x,y)))))
\end{verbatim}
These two versions actually behave differently: the CML version
performs the receives in separate synchronizations, while the TE
version executes both in the \emph{same} synchronization. Therefore
the above TE code could not communicate successfully with code that
performed two synchronizations, such as \ttt{sync (sendEvt c1 4); sync
(sendEvt c2 5)}.
We can mimic \ttt{wrap}'s behavior in TE by thunking the second
receive and executing the thunk after the first \ttt{sync} completes:
\begin{verbatim}
(sync (chooseEvt
(thenEvt (recvEvt c1)
(fun x -> alwaysEvt (fun () -> (x, sync (recvEvt c2)))))
(thenEvt (recvEvt c2)
(fun x -> alwaysEvt (fun () -> (sync (recvEvt c1), x)))))) ()
\end{verbatim}
With the use of two helper functions, the TE code approaches the
elegance of the original CML code:
\begin{verbatim}
(* thunkWrap : 'a event -> ('a -> 'b event) -> (unit -> 'b event) *)
let thunkWrap ev f = thenEvt ev (fun x -> alwaysEvt (fun () -> f x))
(* syncThunked : (unit -> 'a) event -> 'a *)
let syncThunked ev = (sync ev) ()
let _ = syncThunked
(chooseEvt (thunkWrap (recvEvt c1) (fun x -> (x, sync (recvEvt c2))))
(thunkWrap (recvEvt c2) (fun x -> (sync (recvEvt c1), x))))
\end{verbatim}
We have sacrificed some composability: \ttt{thunkWrap} returns a
\ttt{(unit -> 'b) event} instead of a \ttt{'b event}, so it is more
difficult than \ttt{wrap} to combine with other events. The semantics
of \ttt{wrap} (processing an event's result after synchronization
completes) and \ttt{thenEvt} (combining two synchronizations into one)
seem to be incompatible, but we believe that thunked wrappers are a
good compromise. Wrapping an already wrapped event does not require a
second level of thunk, as this helper function demonstrates:
\begin{verbatim}
(* rewrap : (unit -> 'a) event -> ('a -> 'b) -> (unit -> 'b) event *)
let rewrap ev g =
thenEvt ev (fun f -> let x = f () in alwaysEvt (fun () -> g x))
\end{verbatim}
CML's \ttt{guard} is useful for encapsulating actions that need to
happen prior to synchronization. For example, the following code adds
a timeout to an event by spawning a thread to signal when to give up:
\begin{verbatim}
(* timeoutEvt : 'a event -> float -> 'a event *)
let timeoutEvt ev time = guard (fun () ->
let timeoutChan = newChan () in
Thread.create
(fun () -> Thread.delay time; sync (sendEvt timeoutChan ())) ();
chooseEvt ev (wrap (recvEvt timeoutChan) (fun () -> raise TimedOutExn)))
\end{verbatim}
As with \ttt{wrap}, it is difficult to add \ttt{guard} to TE's
interface because the guard function needs to execute outside of the
synchronization. However, we can code up \ttt{timeoutEvt} in TE
without \ttt{guard}:
\begin{verbatim}
let timeoutEvt ev time = chooseEvt ev
(thenEvt (alwaysEvt ()) (fun () -> Thread.delay time; raise TimedOutExn))
\end{verbatim}
Moreover, this function is more readable than the CML code. In the
next section, we will see another example for which the TE
implementation is more natural than the original CML program.
\section{\texttt{wrapAbort} and abort actions}
The \ttt{wrapAbort} function (see Figure \ref{fig:types}) lets CML
programs specify an action to take if an event is part of a
\ttt{chooseEvt} that is synchronized on and another choice is taken.
For example, \ttt{sync (chooseEvt (wrapAbort ev1 f) ev2)} will block
until either \ttt{ev1} or \ttt{ev2} succeeds, and in the latter case
it will execute \ttt{f ()}. \ttt{wrapAbort} is useful for
client-server protocols that have more than one communication. The CML
book \cite{cml-book} uses \ttt{wrapAbort} to implement
mutual-exclusion locks with this interface\footnote{It actually uses
the equally expressive \ttt{withNack}; we discuss \ttt{wrapAbort}
because we find it more intuitive.}:
\begin{verbatim}
type lockServer
val acquireLockEvt : lockServer -> int -> unit event
val releaseLockEvt : lockServer -> int -> unit event
val mkLockServer : unit -> lockServer
\end{verbatim}
Clients of this interface acquire or release locks (represented by
\ttt{int}s) by synchronizing on events created with
\ttt{acquire\-Lock\-Evt} and \ttt{releaseLockEvt}. Synchronizing on
\ttt{acquireLockEvt s i} blocks until the acquire succeeds; clients
may abort acquires, perhaps with the \ttt{timeoutEvt} function from
Section \ref{sec:wrap}:
\begin{verbatim}
sync (timeoutEvt (acquireLockEvt server 1) 5.0)
\end{verbatim}
Implementing \ttt{acquireLockEvt} with a single server thread in CML
requires two communications: a request from the client with the lock
ID, and a confirmation from the server when the lock is available.
Abort actions are essential for implementing \ttt{acquireLockEvt}, as
the first communication may affect the server's internal state by
adding the request to a queue. If the client aborts after the first
communication but before the second, it must tell the server to remove
the request from the queue. Otherwise, the server would hang when
trying to confirm the lock acquire. The form of \ttt{acquireLockEvt}
is essentially:
\begin{verbatim}
let acquireLockEvt s id =
guard (fun () -> (* send acquire request *);
wrapAbort (* receive acquire confirmation *)
(fun () -> (* do cancellation *)))
\end{verbatim}
The abort action effectively makes the two communications
\emph{transactional}: if the second communication aborts, the effects
of the first communication are canceled. In TE, we can implement a
lock server without an abort action. Our solution uses \ttt{thenEvt}
and \ttt{neverEvt} and is similar to the guarded-receive
pattern~\cite{te-hs}. If the server receives a request for an
unavailable lock (the server maintains a list of held locks), it
returns \ttt{neverEvt} inside \ttt{thenEvt}. \ttt{neverEvt} never
succeeds, so the program behaves as if the request did not yet occur,
exploiting the transactional semantics of \ttt{thenEvt}. A full
implementation is below\footnote{As an orthogonal issue, we use
\ttt{serverLoop} from Section 3 to support multiple lock operations
in one synchronization.}; we expect other CML protocols that use
abort actions will adapt similarly to TE\@.
\begin{Verbatim}[commandchars=\\\{\}]
type request = Acquire of int | Release of int
type lockServer = request chan
let acquireLockEvt reqCh id = sendEvt reqCh (Acquire id)
let releaseLockEvt reqCh id = sendEvt reqCh (Release id)
let mkLockServer () =
let reqCh = newChan () in
let serverEvt heldLocks =
\textcolor{blue}{thenEvt} (recvEvt reqCh) (function
| Acquire id -> if List.exists (fun id2 -> id = id2) heldLocks
then \textcolor{blue}{neverEvt}
else alwaysEvt (id::heldLocks)
| Release id -> alwaysEvt (List.filter (fun id2 -> id <> id2) heldLocks))
in Thread.create (serverLoop (serverEvt [], ()))
(fun (heldLocks, ()) -> (serverEvt heldLocks, ()));
reqCh
\end{Verbatim}
\section{Conclusion}
\label{sec:conc}
Every programming model needs three things: a semantics, an
implementation, and useful idioms. Reppy's dissertation on CML
\cite{cml-book} presents all three to demonstrate that CML
is a useful programming model. Prior TE research \cite{te-hs,te-ml}
has concentrated on semantics and implementation. We have presented
several important TE programming idioms, the subtleties of which
surprised us during our research. First, writing client-server
protocols in TE requires careful consideration of how \ttt{sync}
interacts with \ttt{thenEvt}; our \ttt{serverLoop} function is a
general solution to this problem. Second, \ttt{wrap} and \ttt{guard}
are difficult to integrate with TE, and we have suggested alternatives
that preserve most of the original semantics. Finally, we have
discussed how protocols with abort actions may be rewritten
with \ttt{thenEvt} and \ttt{neverEvt}.
\bibliographystyle{eptcs}
|
\section{Introduction}
\label{sec:1}
Neutron stars are natural laboratories for studying the physics of dense
matter. They are associated with some of the most exotic environments in the
universe. Our knowledge of neutron star interiors is still uncertain.
The central density of neutron stars can be extremely high, and many
possibilities for such dense matter have been suggested~\cite{Weber05,PR00,PR07}.
For densities below twice normal nuclear matter density
($\rho_{0} \sim 0.15\;\textrm{fm}^{-3}$), the matter consists of
only nucleons and leptons. When the density is higher than $2 \rho_{0}$,
the equation of state (EOS) and composition of matter are much less certain.
The presence of hyperons in neutron stars has been studied by many
authors~\cite{PRC96,PRC99,Panda04,Shen02,Gl01}. $K^{-}$ condensation in dense
matter was suggested by Kaplan and Nelson~\cite{KN86} and has been extensively discussed
in many works~\cite{PR00,PRC96,YS08}. It has been suggested that the quark
matter may exist in the core of massive neutron stars, and the hadron-quark
phase transition can proceed through a mixed phase of hadronic and quark
matter~\cite{Weber05,PR00,Panda04,PRC08}. If deconfined quark matter does
exist inside stars, it is likely to be in a color superconducting
phase~\cite{Weber05,Panda04}, and various color superconducting phases
have been intensively investigated in recent years~\cite{PR05,Huang,Zhuang}.
Baryon pairing is believed to play an important role in the evolution of
neutron stars~\cite{PR00,RMP03,PRL00}. The presence of neutron superfluidity in
the crust and the inner part of neutron stars can be considered well
established~\cite{RMP03}. The neutron fluid in the crust probably forms
a $^1S_0$ superfluid. With increasing density, the $^1S_0$ interaction turns
repulsive, and the neutrons in the outer core mainly form a $^3P_2$
superfluid. On the other hand, one can expect $^1S_0$ proton pairing in the
outer core because the small proton fraction brings about a low proton
density in this region. In the inner core of neutron stars, hyperons may
appear through the weak interaction because of the fast rise of the baryon
chemical potentials with density. It is widely accepted that hyperons
appear around $2 \rho_{0}$. The presence of hyperons tends to soften
the EOS at high density and lower the maximum mass of neutron
stars~\cite{PRC96,PRC99,Panda04,Shen02,Gl01}
as well as increase the neutron star cooling rate~\cite{PRL00,APJS99}.
In general, the first hyperon to appear is $\Lambda$,
which is the lightest one with an attractive potential in nuclear matter.
The potential of $\Sigma$ hyperons
is now considered to be repulsive; therefore $\Sigma^-$ appears at a higher
density than $\Lambda$ in neutron star matter~\cite{JPG08,YS09}.
The $^1S_0$ superfluidity of $\Lambda$ hyperons is suggested to occur in
the same way as that of neutrons arising from the attractive $\Lambda\Lambda$
interaction in the $^1S_0$ channel~\cite{BB98,TT99,TT00,TMS03,TNYT06}.
It is known that hyperon pairing can significantly affect the thermal
evolution of neutron stars by suppressing neutrino emission from the
hyperon direct Urca process~\cite{TNYT06,APJ98,APJ09}.
Young neutron stars cool primarily by neutrino emission from the interior.
As discussed in Refs.~\cite{Page06,Yakovlev04,Yakovlev01},
the neutrino emissivity in superfluid matter is exponentially suppressed
when the temperature $T$ is much lower than the superfluid critical
temperature $T_c$. On the other hand, superfluidity initiates a specific
neutrino emission from the Cooper pair breaking and formation process,
which is forbidden in nonsuperfluid matter.
This process is exponentially suppressed when $T\ll T_c$, and it is much
less efficient than the direct Urca process~\cite{Yakovlev01,PLB06}.
Hence the presence of baryon superfluidity can
drastically suppress the neutrino emission, which may play a key role in
neutron star cooling. We are mainly interested in the possibility of $^1S_0$
superfluidity of $\Lambda$ hyperons in neutron stars.
So far, the $^1S_0$ pairing gap of $\Lambda$ hyperons is still uncertain
because it can be significantly influenced by both the properties of matter
and the $\Lambda\Lambda$ interaction. More studies are needed to determine
these uncertain factors using available information from recent
developments in hypernuclear physics.
In this article, we focus on the $^1S_0$ superfluidity of $\Lambda$ hyperons
in neutron star matter, which is composed of a chemically equilibrated and
charge-neutral mixture of nucleons, hyperons, and leptons. To calculate the
pairing gap, we need to specify how to treat the neutron star matter and the
$\Lambda\Lambda$ interaction. In this article, we use the relativistic
mean field (RMF) theory to calculate the properties of neutron star matter.
The RMF theory has been successfully and widely used for the description of
nuclear matter and finite nuclei~\cite{Serot86,Ring90,Ren02,Toki96,Shen06}.
It has also been applied to providing the EOS of dense matter for use in
supernovae and neutron stars~\cite{Shen98}. In the RMF approach, baryons
interact through the exchange of scalar and vector mesons. The meson-nucleon
coupling constants are generally determined by fitting to some nuclear
matter properties or ground-state properties of finite nuclei.
To examine the influence of the RMF parameters, we employ two
successful parameter sets, TM1~\cite{ST94} and NL3~\cite{NL3},
which have been widely used in many studies of nuclear
physics~\cite{PRC08,Shen98,Shen06,ST94,NL3,HTT95}.
As for the meson-hyperon couplings, there are large uncertainties because of
limited experimental data in hypernuclear physics. Generally, one can use the
coupling constants derived from the quark model or the values constrained by
reasonable hyperon potentials. The meson-hyperon couplings play an important
role in determining the properties of neutron star matter~\cite{Gl01,Shen03}.
We use the values constrained by reasonable hyperon potentials that
include the updated information from recent developments in hypernuclear
physics. We take into account the two additional hidden-strangeness
mesons, $\sigma^{\ast}$ and $\phi$, which were originally introduced to
obtain the strong attractive $\Lambda\Lambda$ interaction deduced from the
earlier measurement~\cite{PRL93}. A recent observation of the double-$\Lambda$
hypernucleus $_{\Lambda\Lambda}^{6}\textrm{He}$, called the Nagara
event~\cite{Nagara}, has had a significant impact on strangeness nuclear physics.
The Nagara event provides unambiguous identification of
$_{\Lambda\Lambda}^{6}\textrm{He}$ production with a precise $\Lambda\Lambda$
binding energy value
$B_{\Lambda\Lambda }=7.25\pm 0.19_{-0.11}^{+0.18}\;\textrm{MeV}$, which suggests
that the effective $\Lambda\Lambda$ interaction should be considerably
weaker ($\triangle B_{\Lambda\Lambda}\simeq 1\;\textrm{MeV}$) than that
deduced from the earlier measurement
($\triangle B_{\Lambda\Lambda}\simeq$ 4--5 MeV).
The weak hyperon-hyperon ($YY$) interaction suggested by the Nagara event
has been used to reinvestigate the properties of multistrange systems,
and it has been found that the change of $YY$ interactions affects
the properties of strange hadronic matter
dramatically~\cite{PRC08,PRC03s,JPG04s,JPG05}.
We would like to examine whether the $^1S_0 $ superfluidity of $\Lambda$
hyperons exists in neutron star matter, and how large the pairing gap
can be if it does, by considering recent developments in hypernuclear physics.
The aim of this article is to investigate the possibility of forming
a $\Lambda$ superfluid in neutron stars. It has been suggested
that hyperon superfluidity could significantly suppress the neutrino emission
in the core of a neutron star and play a key role in neutron star
cooling~\cite{TNYT06,APJ98}.
Over the last decade, there has been some discussion in the literature
about hyperon pairing in dense matter~\cite{BB98,TT99,TT00,TMS03,TNYT06,SIGMA04}.
In the work of Balberg and Barnea~\cite{BB98}, the $^1S_0$ superfluidity of $\Lambda$
hyperons has been studied by using an effective $\Lambda\Lambda$ interaction
based on a $G$ matrix calculation and an approximation of
nonrelativistic effective mass obtained from single-particle energies.
Their calculation predicts a gap energy of a few tenths of a MeV
for $\Lambda$ Fermi momenta up to about $1.3$ $\textrm{fm}^{-1}$.
In Refs.~\cite{TT99,TT00}, Takatsuka and Tamagaki studied $\Lambda$ superfluidity
using two types of bare $\Lambda\Lambda$ interactions based on the
one-boson-exchange (OBE) model and two types of hyperon core models.
They found that $\Lambda$ superfluidity could exist in a density region
between 2$\rho_{0}$ and (2.6--4.6)$\rho_{0}$, depending on the pairing
interaction and the hyperon core model.
A study of $\Lambda\Lambda$ pairing in a pure neutron background
has been presented by Tanigawa \emph{et al.}~\cite{TMS03} using the
relativistic Hartree-Bogoliubov model, where the $\Lambda\Lambda$ pairing gap
was found to decrease with increasing background density
and decreasing $\Lambda\Lambda$ attraction.
Both $\Lambda$ and $\Sigma^-$ superfluidities in neutron star matter have been
investigated by Takatsuka \emph{et al.}~\cite{TNYT06} using three pairing
interactions based on the OBE model and several nonrelativistic EOS with different
incompressibilities. It was found that both $\Lambda$ and $\Sigma^-$ are
superfluid as soon as they begin to appear at around $4 \rho_{0}$,
although the pairing gap and the density region depend on the pairing
interaction and the EOS of neutron star matter.
The effect of the Nagara event on $\Lambda$ superfluidity has been discussed
in Refs.~\cite{TMS03,TNYT06}, where the weak attractive $\Lambda\Lambda$
interaction suggested by the Nagara event leads to very small pairing
gap in dense neutron matter~\cite{TMS03} or the disappearance of
$\Lambda$ superfluidity in neutron star matter~\cite{TNYT06}.
All these studies indicate that the $\Lambda\Lambda$ pairing gap
in dense matter depends both on the $\Lambda\Lambda$ interaction
and on the EOS of matter. The key role of $\Lambda$ superfluidity
in neutron star cooling motivates us to investigate the possibility of
forming a $\Lambda$ superfluid in neutron stars by carefully considering
the pairing interaction and the description of neutron star matter
with the updated information from recent developments in
hypernuclear physics.
This article is arranged as follows. In Sec.~\ref{sec:2}, we briefly
describe the RMF theory for the calculation of neutron star matter
properties. In Sec.~\ref{sec:3}, we discuss the $\Lambda\Lambda$
interaction used in the gap equation. We present the numerical results
in Sec.~\ref{sec:4}. Section~\ref{sec:5} is devoted to a summary.
\section{ Relativistic mean field theory}
\label{sec:2}
We use the RMF theory to describe the neutron star matter, which is composed
of a chemically equilibrated and charge-neutral mixture of nucleons,
hyperons, and leptons. In the RMF approach, baryons interact through the
exchange of scalar and vector mesons. The baryons considered in this work
are nucleons ($p$ and $n$) and hyperons ($\Lambda$, $\Sigma$, and $\Xi$).
The exchanged mesons include isoscalar scalar and vector mesons
($\sigma$ and $\omega$), an isovector vector meson ($\rho$), and two additional
hidden-strangeness mesons ($\sigma^{\ast }$ and $\phi$).
The total Lagrangian density of neutron star matter takes the form
\begin{eqnarray}
\mathcal{L}_{RMF} &=&\sum_{B}\bar{\psi}_{B}\left[ i\gamma_{\mu}
\partial^{\mu}-m_{B}-g_{\sigma B}\sigma-g_{\sigma^{\ast}B} \sigma^{\ast}
-g_{\omega B}\gamma_{\mu}\omega^{\mu}\right. \notag \\
&&\left. -g_{\phi B}\gamma_{\mu}\phi^{\mu}-g_{\rho B}\gamma_{\mu}
\tau_{iB}\rho_{i}^{\mu}\right] \psi_{B} +\frac{1}{2}\partial_{\mu}
\sigma\partial^{\mu}\sigma -\frac{1}{2}m_{\sigma}^{2}\sigma^{2} \notag \\
&&-\frac{1}{3}g_{2}\sigma^{3}-\frac{1}{4}g_{3}\sigma^{4}
-\frac{1}{4}W_{\mu\nu}W^{\mu\nu}+\frac{1}{2}m_{\omega}^{2}
\omega_{\mu}\omega^{\mu} \notag \\
&&+\frac{1}{4}c_{3}\left( \omega_{\mu}\omega^{\mu}\right)^{2}
-\frac{1}{4}R_{i\mu\nu}R_{i}^{\mu\nu}+\frac{1}{2}m_{\rho}^{2}
\rho_{i\mu}\rho_{i}^{\mu} \notag \\
&&+\frac{1}{2}\partial_{\mu}\sigma^{\ast}\partial^{\mu}\sigma^{\ast}
-\frac{1}{2}m_{\sigma^{\ast}}^{2}\sigma^{\ast 2}-\frac{1}{4} S_{\mu\nu}S^{\mu\nu}
+\frac{1}{2}m_{\phi}^{2}\phi_{\mu}\phi^{\mu} \notag \\
&&+\sum_{l}\bar{\psi}_{l}\left[ i\gamma_{\mu}\partial^{\mu}-m_{l}\right]
\psi_{l},
\label{eq:Lrmf}
\end{eqnarray}
where $\psi_{B}$ and $\psi_{l}$ are the baryon and lepton fields,
respectively. The index $B$ runs over the baryon octet
($p$, $n$, $\Lambda$, $\Sigma^{+}$, $\Sigma^{0}$, $\Sigma^{-}$,
$\Xi^{0}$, $\Xi^{-}$), and the sum on $l$ is over electrons and muons
($e^{-}$ and $\mu^{-}$). The field tensors of the vector mesons,
$\omega$, $\rho$, and $\phi$, are denoted by $W_{\mu\nu}$, $R_{i\mu\nu}$,
and $S_{\mu\nu}$, respectively. In the RMF approach, the meson fields are
treated as classical fields, and the field operators are replaced by their
expectation values. The meson field equations in uniform matter have the
following form:
\begin{eqnarray}
&&m_{\sigma }^{2}\sigma +g_{2}\sigma ^{2}+g_{3}\sigma ^{3}
=-\sum_{B}\frac{g_{\sigma B}}{\pi ^{2}}\int_{0}^{k_{F}^{B}}
\frac{m_{B}^{\ast }}{\sqrt{k^{2}+m_{B}^{\ast 2}}}k^{2}dk,
\label{eq:s} \\
&&m_{\omega }^{2}\omega +c_{3}\omega ^{3}=\sum_{B}\frac{g_{\omega B}
\left(k_{F}^{B}\right)^{3}}{3\pi ^{2}},
\label{eq:w} \\
&&m_{\rho }^{2}\rho =\sum_{B}\frac{g_{\rho B}\tau _{3B}
\left(k_{F}^{B}\right) ^{3}}{3\pi ^{2}},
\label{eq:r} \\
&&m_{\sigma ^{\ast }}^{2}\sigma ^{\ast }=-\sum_{B}
\frac{g_{\sigma ^{\ast }B}}{\pi ^{2}}\int_{0}^{k_{F}^{B}}
\frac{m_{B}^{\ast }}{\sqrt{k^{2}+m_{B}^{\ast 2}}}k^{2}dk,
\label{eq:ss} \\
&&m_{\phi }^{2}\phi =\sum_{B}\frac{g_{\phi B}\left( k_{F}^{B}\right)^{3}}
{3\pi ^{2}},
\label{eq:ws}
\end{eqnarray}
where $\sigma =\left\langle \sigma \right\rangle$,
$\omega =\left\langle \omega^{0}\right\rangle$,
$\rho =\left\langle \rho_{30}\right\rangle$,
$\sigma^{\ast} =\left\langle \sigma^{\ast}\right\rangle$,
and $\phi=\left\langle \phi^{0}\right\rangle$
are the nonvanishing expectation values of meson fields in uniform matter;
$m_{B}^{\ast}=m_{B}+g_{\sigma B}\sigma+g_{\sigma^{\ast}B}\sigma^{\ast}$
is the effective mass of the baryon species $B$, and $k_{F}^{B}$ is the
corresponding Fermi momentum.
The meson-baryon coupling constants play an important role in determining
the properties of neutron star matter.
To examine the influence of the RMF parameters, we employ two
successful parameter sets, TM1~\cite{ST94} and NL3~\cite{NL3},
in the present calculation.
These parameters have been determined by fitting to some ground-state
properties of finite nuclei, and they can provide a good description of
nuclear matter and finite nuclei, including unstable nuclei.
With the TM1 (NL3) parameter set, the nuclear matter saturation
density is $0.145$ fm$^{-3}$ ($0.148$ fm$^{-3}$),
the energy per nucleon is $-16.3$ MeV ($-16.3$ MeV),
the symmetry energy is $36.9$ MeV ($37.4$ MeV),
and the incompressibility is $281$ MeV ($272$ MeV)~\cite{ST94,NL3}.
As for the meson-hyperon couplings, we take the
naive quark model values for the vector coupling constants;
\begin{eqnarray}
&&\frac{1}{3}g_{\omega N}=\frac{1}{2}g_{\omega \Lambda }=\frac{1}{2}
g_{\omega \Sigma }=g_{\omega \Xi }, \notag \\
&&g_{\rho N}=\frac{1}{2}g_{\rho \Sigma }=g_{\rho \Xi },
\ \ g_{\rho\Lambda}=0, \notag \\
&&2g_{\phi \Lambda }=2g_{\phi \Sigma }=g_{\phi \Xi }=-\frac{2\sqrt{2}}{3}
g_{\omega N},\ \ g_{\phi N}=0.
\end{eqnarray}
The scalar coupling constants are chosen to give reasonable hyperon
potentials. We denote the potential depth of the hyperon species $i$ in
the matter of the baryon species $j$ by $U_{i}^{\left( j\right) }$.
It is estimated from the experimental data of single-$\Lambda$ hypernuclei
that the potential depth of a $\Lambda$ in saturated nuclear matter should
be around $U_{\Lambda}^{\left(N\right)} \simeq -30$ MeV~\cite{PRC00}.
For $\Sigma$ hyperons, the analysis of $\Sigma$ atomic experimental data
suggests that $\Sigma$-nucleus potentials have a repulsion inside the nuclear
surface and an attraction outside the nucleus with a sizable absorption.
In recent theoretical works, the $\Sigma$ potential in saturated nuclear
matter is considered to be repulsive with a strength of about
$30$ MeV~\cite{JPG08,PRC00}. Some recent developments in hypernuclear physics
suggest that $\Xi$ hyperons in saturated nuclear matter have an attraction
of around $15$ MeV~\cite{JPG08,PRC00xi}. In this article, we adopt
$U_{\Lambda }^{\left(N\right)}=-30$ MeV,
$U_{\Sigma }^{\left(N\right) }=+30$ MeV, and
$U_{\Xi}^{\left(N\right)}=-15$ MeV
to determine the scalar coupling constants.
We obtain, for the TM1 (NL3) parameter set,
$g_{\sigma\Lambda}=6.228$ (6.323),
$g_{\sigma\Sigma}=4.472$ (4.709), and
$g_{\sigma\Xi}=3.114$ (3.161), respectively.
The hyperon couplings to the hidden-strangeness meson
$\sigma^{\ast}$ are restricted by the relation
$U_{\Xi }^{\left(\Xi\right)} \approx
U_{\Lambda}^{\left(\Xi\right) } \approx
2U_{\Xi}^{\left(\Lambda\right) } \approx
2U_{\Lambda}^{\left(\Lambda\right) }$ obtained in Ref.~\cite{ANN94}.
The weak $YY$ interaction implied by the Nagara event suggests
$U_{\Lambda}^{(\Lambda)}\simeq -5$ MeV, and hence we obtain
$g_{\sigma^*\Lambda}=5.499$ (5.678)
and $g_{\sigma^*\Xi}=11.655$ (11.899) for the TM1 (NL3) parameter set.
We assume $g_{\sigma^*\Sigma}=g_{\sigma^*\Lambda}$
and take $m_{\sigma^{\ast }}=980$ MeV and $m_{\phi }=1020$
MeV in this article.
For neutron star matter consisting of a neutral mixture of baryons and
leptons, the $\beta$-equilibrium conditions without trapped neutrinos
are given by
\begin{eqnarray}
&&\mu_{p}=\mu_{\Sigma^{+}}=\mu_{n}-\mu_{e}, \label{eq:beta1} \\
&&\mu_{\Lambda}=\mu_{\Sigma^{0}}=\mu_{\Xi^{0}}=\mu_{n}, \label{eq:beta2} \\
&&\mu_{\Sigma^{-}}=\mu_{\Xi^{-}}=\mu_{n}+\mu_{e}, \label{eq:beta3} \\
&&\mu_{\mu}=\mu_{e}, \label{eq:beta4}
\end{eqnarray}
where $\mu_{i}$ is the chemical potential of species $i$.
At zero temperature, the chemical potentials of baryons and leptons are
given by
\begin{eqnarray}
&&\mu_{B}=\sqrt{{k_{F}^{B}}^{2}+m_{B}^{\ast 2}}+g_{\omega B}\omega
+g_{\phi B}\phi +g_{\rho B}\tau_{3B}\rho , \label{eq:mub} \\
&&\mu_{l}=\sqrt{{k_{F}^{l}}^{2}+m_{l}^{2}}, \label{eq:mul}
\end{eqnarray}
respectively. The electric charge neutrality condition is expressed by
\begin{eqnarray}
\rho_{p}+\rho_{\Sigma^{+}} =\rho_{e}+\rho_{\mu}
+\rho_{\Sigma^{-}}+\rho_{\Xi^{-}}, \label{eq:charge}
\end{eqnarray}
where $\rho_{i}=\left( k_{F}^{i}\right)^{3}/(3\pi^{2})$ is the number
density of species $i$. We solve the coupled
Eqs.~(\ref{eq:s})--(\ref{eq:ws}), (\ref{eq:beta1})--(\ref{eq:beta4}),
and (\ref{eq:charge}) self-consistently at a given baryon density $\rho_{B}$.
Then we can calculate the EOS and the composition of neutron star matter
as well as the effective mass and the Fermi momentum of $\Lambda$ hyperons,
which are crucial in the study of $\Lambda\Lambda$ pairing.
\section{ Gap equation and $\Lambda\Lambda$ interaction}
\label{sec:3}
We study the $^{1}S_{0}$ superfluidity of $\Lambda $ hyperons in neutron
star matter. The crucial quantity in determining the onset of superfluidity
is the energy gap function $\Delta \left( k\right) $, which can be obtained
by solving the gap equation
\begin{equation}
\Delta \left( k\right) =-\frac{1}{4\pi ^{2}}\int k^{\prime 2}dk^{\prime }
\frac{V\left( k,k^{\prime }\right) \Delta \left( k^{\prime }\right) }
{\sqrt{\left[ E\left( k^{\prime }\right) -E\left( k_{F}^{\Lambda }\right)
\right]^{2}+\Delta ^{2}\left( k^{\prime }\right) }},
\label{gapeq}
\end{equation}
where $E\left( k\right) $ is the single-particle energy of $\Lambda$
with momentum $k$. For $\Lambda $ hyperons in neutron star matter,
the single-particle energy in the RMF approach is given by
\begin{equation*}
E\left( k\right) =\sqrt{{k}^{2}+m_{\Lambda }^{\ast 2}}
+g_{\omega\Lambda}\omega +g_{\phi\Lambda}\phi .
\end{equation*}
The effective mass $m_{\Lambda }^{\ast }$ and the Fermi momentum
$k_{F}^{\Lambda }$ are computed self-consistently at a given
baryon density $\rho_{B}$ within the RMF approach.
For the $\Lambda\Lambda$ pairing interaction in the $^{1}S_{0}$
channel, the potential matrix element can be written as
\begin{equation}
V\left( k,k^{\prime }\right) =\left\langle k\right\vert
V_{\Lambda \Lambda}\left( ^{1}S_{0}\right) \left\vert k^{\prime }\right\rangle
=4\pi \int r^{2}dr\text{ }j_{0}(kr)V_{\Lambda \Lambda }
\left( r\right) j_{0}(k^{\prime}r),
\end{equation}
where $j_{0}(kr)=\textrm{sin}(kr)/(kr)$ is the spherical Bessel function
of order zero and $V_{\Lambda \Lambda }\left( r\right) $ is the
$^{1}S_{0}$ $\Lambda\Lambda $ interaction potential in coordinate space.
Because of large uncertainties in the $\Lambda\Lambda$ interaction,
we adopt several $\Lambda\Lambda$ potentials. Most of them are based on
the Nijmegen models and are used in double-$\Lambda$ hypernuclei studies,
which are of the three-range Gaussian form
\begin{equation}
V_{\Lambda \Lambda }(r)=\sum_{i=1}^{3}v_{i}\exp (-r^{2}/\beta_{i}^{2}).
\label{eq:vll}
\end{equation}
The short-range term provides for a strong soft-core repulsion, whereas the
medium-range and long-range terms provide for attraction. The parameters $v_{i}$
and $\beta_{i}$ are taken from Refs.~\cite{Hiy97,NPA02,Hiy02,PRL02}, and we
list them in Table~\ref{tab:1}. The ND1 potential was given in Ref.~\cite{Hiy97}
as an effective soft-core interaction fitted to the Nijmegen model D (ND)
hard-core interaction. Another simulation of the ND interaction, called ND2 in
this article, was obtained in Ref.~\cite{NPA02}. The ESC00, NSC97b, NSC97e,
and NSC97f potentials given in Ref.~\cite{NPA02} were obtained by changing
the strength of the medium-range attractive component of the three-range
Gaussian potential such that they could reproduce the scattering length and
the effective range as close to values by the corresponding Nijmegen models.
It is well known that the Nagara event provides unambiguous identification
of $_{\Lambda \Lambda }^{6}\textrm{He}$ production with a precise
$\Lambda\Lambda$ binding energy value $B_{\Lambda\Lambda }$ and has had a
significant impact on strangeness nuclear physics. The NFs and NSC97s
potentials given in Refs.~\cite{Hiy02,PRL02} were obtained by adjusting
parameters to reproduce the experimental value of
$B_{\Lambda\Lambda}(_{\Lambda\Lambda}^{6}\textrm{He})$ from the Nagara event.
We have also chosen an Urbana-type potential that has
been successfully used to explain the experimental values of hypernuclei.
The Urbana potential could be found in Ref.~\cite{PRC06}.
We plot in Fig.~\ref{fig:VR} all $\Lambda\Lambda$ potentials
considered in the present work. The strongest $\Lambda\Lambda$ interaction
is the ESC00 potential, whereas the weakest $\Lambda\Lambda$ interaction
is the NSC97f potential.
We note that the NFs, NSC97s, and Urbana potentials simulate the experimental
value of $B_{\Lambda\Lambda}(_{\Lambda\Lambda}^{6}\textrm{He})$
from the Nagara event.
\section{ Results and discussion}
\label{sec:4}
In this section, we investigate the $^1S_0 $ superfluidity of
$\Lambda$ hyperons in neutron star matter and neutron stars.
We employ the RMF model with the parameter sets TM1 and NL3
to calculate the properties of neutron star matter, which is known
to provide excellent descriptions of the ground states of finite
nuclei, including unstable nuclei.
The meson-hyperon couplings play an important role in determining
the properties of neutron star matter. We use the values constrained
by reasonable hyperon potentials that include the updated information
from recent developments in hypernuclear physics.
As for the $\Lambda\Lambda$ pairing interaction used in the gap equation,
we adopt several $\Lambda\Lambda$ potentials that have been used in
double-$\Lambda$ hypernuclei studies. Some simulate the
experimental value of $B_{\Lambda\Lambda}(_{\Lambda\Lambda}^{6}\textrm{He})$
from the Nagara event. With the effective mass and the Fermi momentum
of $\Lambda$ hyperons obtained in the RMF approach, the gap
equation [Eq.~(\ref{gapeq})] is solved numerically.
In Fig.~\ref{fig:RES}, we show the resulting $^1S_0$ pairing gap of
$\Lambda$ hyperons at the Fermi surface, $\Delta_{F}$, as a function of
the baryon density, $\rho_{B}$, in neutron star matter. The results of TM1 and
NL3 are plotted in Fig.~\ref{fig:RES} (top) and Fig.~\ref{fig:RES} (bottom),
respectively. In the case of TM1 (NL3), the threshold density of $\Lambda$
is around $0.31\;\textrm{fm}^{-3}$ ($0.28\;\textrm{fm}^{-3}$),
and $\Lambda$ hyperons form a $^1S_0$ superfluid as soon as they appear
in neutron star matter. With increasing baryon density, $\Delta_{F}$ increases
first, reaching a maximum value at $\rho_{B} \sim 0.34\;\textrm{fm}^{-3}$
($\rho_{B} \sim 0.30\;\textrm{fm}^{-3}$), then decreases and finally vanishes
at $\rho_{B}< 0.46\;\textrm{fm}^{-3}$ ($\rho_{B}< 0.38\;\textrm{fm}^{-3}$)
for the case of TM1 (NL3) with the ESC00 potential.
It is found that the maximal pairing gap is about 0.8 MeV with the
ESC00 potential in the TM1 case. This is because the ESC00 potential
has the strongest attraction among the $\Lambda\Lambda$ interactions used here.
The pairing gaps with the ND1 and ND2 potentials are of the order of
$0.1$--$0.2$ MeV, as shown in Fig.~\ref{fig:RES}. In addition, we find that the
pairing gaps are of the order of $10^{-4}$ MeV (TM1) or absent (NL3)
with the NSC97e, NFs, NSC97s, and Urbana potentials.
The $\Lambda$ pairing does not appear for the NSC97b and NSC97f potentials.
We present in Table~\ref{tab:2} the maximal pairing gap at the Fermi
surface ($\Delta_F^{\textrm{max}}$)
and the corresponding baryon density ($\rho_B$), effective $\Lambda$
mass ($m_\Lambda^\ast$), and Fermi momentum ($k_F^\Lambda$)
using these potentials with the TM1 and NL3 parameter sets.
The $\Lambda$ pairing gap $\Delta_{F}$ depends not only on the $\Lambda\Lambda$
interaction but also on the properties of $\Lambda$ hyperons in neutron
star matter. In Fig.~\ref{fig:YI}, we show the particle fraction,
$Y_i=\rho_i/\rho_B$, as a function of the baryon density, $\rho_{B}$, using the
RMF model with the TM1 (Fig.~\ref{fig:YI}, top) and NL3 (Fig.~\ref{fig:YI}, bottom)
parameter sets.
It is seen that $\Lambda$ hyperons appear around $0.31\;\textrm{fm}^{-3}$ (TM1)
or $0.28\;\textrm{fm}^{-3}$ (NL3) and then increase rapidly with increasing density.
We note that hyperon threshold densities, fractions, and effective masses are
dependent on the RMF parameters used. This dependence has an effect on
the resulting pairing gap, as shown in Fig.~\ref{fig:RES}.
Our results with the ND1 potential can be compared with those in Table III
of Ref.~\cite{TNYT06}, where the same $\Lambda\Lambda$ interaction
(called the ND-Soft potential) was used. The difference is in the treatment
of neutron star matter, for which they use a nonrelativistic $G$ matrix-based
effective interaction approach, whereas we use the RMF approach.
In our case of TM1 (NL3), the maximal pairing gap at the Fermi surface is
0.17 MeV (0.12 MeV), as given in Table~\ref{tab:2},
where $\rho_B=0.344\;\textrm{fm}^{-3}$ ($\rho_B=0.303\;\textrm{fm}^{-3}$),
$Y_\Lambda=0.039$ ($Y_\Lambda=0.044$), and
$m_{\Lambda}^{\ast}=743$ MeV ($m_{\Lambda}^{\ast}=706$ MeV).
Takatsuka \textit{et al}.~\cite{TNYT06} obtained the maximal pairing gap of 0.34 MeV at
$\rho_B=4.5\rho_{0}$ for the TNI6u EOS. The larger pairing gap at higher
$\rho_B$ given in Ref.~\cite{TNYT06} is because of their smaller $Y_\Lambda$
and larger $m_{\Lambda}^{\ast}$.
As discussed in Refs.~\cite{BB98,TMS03,TNYT06}, the pairing gap is very
sensitive to the effective mass. Generally, a smaller effective mass leads to a
higher single-particle energy and then yields a smaller pairing gap.
In Fig.~\ref{fig:MSTAR}, we show the effective mass of $\Lambda$ hyperons,
$m_{\Lambda}^{\ast}$, as a function of the baryon density, $\rho_{B}$,
using the RMF model with the TM1 (solid line) and NL3 (dashed line) parameter sets.
It is shown that $m_{\Lambda}^{\ast}$ decreases with increasing $\rho_{B}$.
When $\Lambda$ hyperons appear around $0.31\;\textrm{fm}^{-3}$ (TM1)
or $0.28\;\textrm{fm}^{-3}$ (NL3), the effective mass of $\Lambda$ hyperons
is about 762 MeV or 727 MeV. It is found that the effective masses of $\Lambda$
hyperons in the NL3 case are smaller than those of TM1, which leads to a
smaller pairing gap, as shown in Fig.~\ref{fig:RES}.
We note that the effective mass of $\Lambda$ hyperons is mainly determined by
the coupling constants $g_{\sigma\Lambda}$ and $g_{\sigma^{\ast}\Lambda}$.
Here we use the values constrained by reasonable hyperon potentials,
which are suggested by the experimental data of single-$\Lambda$ hypernuclei
and by the Nagara event.
To examine whether the $^1S_0 $ superfluidity of $\Lambda$ hyperons exists
in neutron stars, we solve the Tolman-Oppenheimer-Volkoff (TOV) equation
with the EOS of the RMF theory over a wide density range.
For the nonuniform matter at low density, which exists in the inner and
outer crusts of neutron stars, we adopt a relativistic EOS based on the
RMF theory with a local density approximation~\cite{Shen02,Shen98}.
The nonuniform matter is modeled to be composed of a lattice of
spherical nuclei immersed in an electron gas with or without
free neutrons dripping out of nuclei. The low-density EOS is
matched to the EOS of uniform matter at the density
where they have equal pressure. The neutron star properties
are mainly determined by the EOS at high density.
Using the EOS described in Sec.~\ref{sec:2}, we calculate the neutron
star properties and find that the maximum mass of neutron stars
is about 1.70 $M_{\odot}$ (2.06 $M_{\odot}$)
with the TM1 (NL3) parameter set.
According to the compilation of measured neutron star masses~\cite{PR07,NS09},
some massive neutron stars were reported to be observed recently.
However, the uncertainties in these mass measurements are rather large,
and the mass of PSR J0751$+$1807 was corrected from
$(2.1 \pm 0.2)$ $M_{\odot}$ to $(1.26 \pm 0.14)$ $M_{\odot}$~\cite{NS08}.
We note that the EOS used here could not be ruled out by current observations.
In Figs.~\ref{fig:allT} and~\ref{fig:allN},
we show the central baryon density as a function of the neutron star mass.
We find that whether the $^1S_0 $ superfluidity of $\Lambda$ hyperons
exists in the core of neutron stars depends on the $\Lambda\Lambda$
interaction used. With weaker $\Lambda\Lambda$ interactions, such as
NSC97b and NSC97f, the $\Lambda$ superfluidity does not appear inside neutron
stars. For the NSC97e, NFs, NSC97s, and Urbana interactions,
although we obtain the pairing gaps of the order of $10^{-4}$ MeV in
the TM1 case, it is unlikely that $\Lambda$ superfluidity can exist
in observed neutron stars because of its low superfluid critical temperature
$T_c \simeq 0.57 \Delta_{F} / \kappa_B \sim 10^6$ K~\cite{TNYT06,Yakovlev01}.
With stronger $\Lambda\Lambda$ interactions, such as ESC00, ND1, and ND2,
the $^1S_0 $ superfluidity of $\Lambda$ hyperons may exist
in massive neutron stars, as shown in Figs.~\ref{fig:allT} and~\ref{fig:allN}.
In the case of TM1 (NL3) with the ESC00 potential,
$\Lambda$ hyperons do not appear in neutron stars
with $M < 1.37 \ M_{\odot}$ ($M < 1.50 \ M_{\odot}$).
For neutron stars with $1.37 \ M_{\odot} < M < 1.63 \ M_{\odot}$
($1.50 \ M_{\odot} < M < 1.82 \ M_{\odot}$), $\Lambda$ hyperons
in the core of neutron stars form a $^1S_0$ superfluid.
However, when $M > 1.63 \ M_{\odot}$ ($M > 1.82 \ M_{\odot}$),
not only superfluid $\Lambda$ but also normal (nonsuperfluid)
$\Lambda$ can exist in the core of neutron stars
because the central baryon density exceeds the upper limit
of the range where $\Lambda$ superfluidity exists.
The presence of nonsuperfluid $\Lambda$ hyperons in the core of
massive stars would lead to a more rapid cooling than the case with
only superfluid $\Lambda$ hyperons.
The mass region, where only superfluid $\Lambda$ hyperons
exist in the core of neutron stars, is shaded
in Figs.~\ref{fig:allT} and~\ref{fig:allN}.
It is shown that the region with the ESC00 potential is the widest
among all cases in these figures. This is because the ESC00 potential
has the strongest attraction, and its pairing gap covers the widest
density range, as shown in Fig.~\ref{fig:RES}.
We note that this region depends both on the $\Lambda\Lambda$
interaction and on the EOS of neutron star matter.
\section{Summary}
\label{sec:5}
We have studied the $^1S_0$ superfluidity of $\Lambda$ hyperons
in neutron star matter and neutron stars. In this article,
we employ the RMF model with the parameter sets TM1 and NL3
to calculate the properties of neutron star matter,
which is composed of a chemically equilibrated and
charge-neutral mixture of nucleons, hyperons, and leptons.
The RMF theory has been successfully and
widely used for the description of nuclear matter and finite nuclei,
including unstable nuclei. In the RMF approach, baryons interact
through the exchange of scalar and vector mesons.
The baryons considered in this article are nucleons ($p$ and $n$)
and hyperons ($\Lambda$, $\Sigma$, and $\Xi$).
The exchanged mesons include isoscalar scalar and vector mesons
($\sigma$ and $\omega$), an isovector vector meson ($\rho$), and two
additional hidden-strangeness mesons ($\sigma^{\ast }$ and $\phi$).
It is well known that the meson-hyperon couplings play an important
role in determining the properties of neutron star matter.
We have used the couplings constrained by reasonable hyperon
potentials that include the updated information from recent
developments in hypernuclear physics.
To examine the $^1S_0$ pairing of $\Lambda$ hyperons, we have adopted
several $\Lambda\Lambda$ potentials. Most are based on the
Nijmegen models and have been used in double-$\Lambda$ hypernuclei studies.
NFs, NSC97s, and Urbana potentials have simulated the experimental
value of $B_{\Lambda\Lambda}(_{\Lambda\Lambda}^{6}\textrm{He})$
from the Nagara event.
We have calculated the $^1S_0$ pairing gap of $\Lambda$ hyperons at the
Fermi surface, $\Delta_{F}$, using the $\Lambda\Lambda$ potentials
adopted in this article. It is found that $\Delta_{F}$ depends
both on the $\Lambda\Lambda$ interaction and on the treatment
of neutron star matter. The maximal $\Delta_{F}$ obtained in
the present calculation is about 0.8 MeV with the ESC00 potential in the TM1 case.
This is because the ESC00 potential has the strongest attraction
among the $\Lambda\Lambda$ interactions used in this article.
The ND1 and ND2 potentials yield somewhat smaller $\Delta_{F}$ of the order
of $0.1$--$0.2$ MeV. For the NSC97e, NFs, NSC97s, and Urbana potentials,
the values of $\Delta_{F}$ are of the order of $10^{-4}$ MeV (TM1) or absent (NL3).
The $\Lambda$ pairing does not appear for the NSC97b and NSC97f potentials.
The difference in these results reflects the dependence of $\Delta_{F}$
on the $\Lambda\Lambda$ interaction.
On the other hand, the magnitude and the threshold density of $\Delta_{F}$
are also dependent on properties of neutron star matter, especially
on the effective mass and particle fraction of $\Lambda$ hyperons.
In the case of TM1 (NL3) with the ESC00 potential,
the threshold density of $\Delta_{F}$ is around
$0.31\;\textrm{fm}^{-3}$ ($0.28\;\textrm{fm}^{-3}$),
reaches a maximum value at $\rho_{B} \sim 0.34\;\textrm{fm}^{-3}$
($\rho_{B} \sim 0.30\;\textrm{fm}^{-3}$), and finally vanishes at
$\rho_{B}< 0.46\;\textrm{fm}^{-3}$ ($\rho_{B}< 0.38\;\textrm{fm}^{-3}$).
By solving the TOV equation, we have calculated neutron star properties
and found that whether the $^1S_0 $ superfluidity of $\Lambda$ hyperons
exists in the core of neutron stars mainly depends on the $\Lambda\Lambda$
interaction used. With stronger $\Lambda\Lambda$ interactions, such as
ESC00, ND1, and ND2, the $\Lambda$ superfluidity may exist in massive neutron
stars. It is unlikely that $\Lambda$ superfluidity can exist
in neutron stars with the NFs, NSC97s, and Urbana interactions,
which have simulated the experimental value of
$B_{\Lambda\Lambda}(_{\Lambda\Lambda}^{6}\textrm{He})$
from the Nagara event.
In this article, we have considered the updated information from recent
developments in hypernuclear physics and used the weak attractive
$\Lambda\Lambda$ interactions suggested by the Nagara event.
However, there are still large uncertainties in the hyperon-hyperon
interaction and the EOS of neutron star matter.
A more precise study of the $\Lambda$ pairing in neutron stars
requires further development in hypernuclear physics.
\section*{ACKNOWLEDGMENT}
This work was supported in part by the National Natural Science
Foundation of China (Grant No. 10675064).
|
\section{Introduction}
A \emph{flock} consists of a large number of moving physical objects, called \emph{agents}, with their positions being controlled in such a way that they move along a prescribed path, in a prescribed and fixed configuration (or \emph{in formation}). Each agent knows the position and velocity of only a few other agents, and this flux of information defines a \emph{communication graph}.
In \cite{flocks2} graph theory and linear systems techniques are combined to provide a framework for studying the control of formations. The main tools of graph theory that are related to the problem are the directed graph Laplacians and the connectedness of the graph. Linear feedback is then used to stabilize the patterns.
More recently \cite{flocks4} studied a system of coupled linear differential equations describing the movement of cars in $\mathbb{R}$, where each car reacts only to its immediate neighbors, and only the movement of the first agent (the \emph{leader}) is independent from the rest of the group. In the paper it was
proved that When equal attention is paid to both neighbors perturbations in the orbit of the leader grow as they propagate through the flock. In fact, perturbations grow proportional to size of the flock: when the leader's perturbation has amplitude 1, then the perturbation in the orbit of the
agent furthest away from the leader will be proportional to $N$ (the size of the flock).
The aim of this paper is first of all to study the (asymptotic) stability of a family
of such systems. This is answered in detail in Theorems \ref{theo:stable}
and \ref{theo:stable2}. Next we assume the flock has a leader that chooses its
orbit independently of the other members of the flock. We then analyze how exactly
the system converges to a stable flight pattern if the leader changes its orbit.
This question only makes sense when the system is already asymptotically stable, which is therefore
assumed henceforth. The latter question is important in applications as too great fluctuations
in the course of convergence to a coherent flight pattern will make that flock unviable.
In all of these arguments we closely follow the reasoning set forth in previous works \cite{flocks5,flocks6,flocks7}. However there are two important differences. The first is that
the farthest member of the flock (in this work) is coupled to the leader. In the language
of partial differential equations, this is akin to changing a boundary condition. The reason is twofold.
Changing the boundary condition can greatly aid the mathematics, and therefore help
to gain insight. The second reason is a deeper one: we do not know how these boundary
condition influence the stability of these systems, and thus this note can be viewed
a test case (when compared with the papers just cited). The other difference with
the previous papers is that we here allow the weight of the coupling with neighbors
to be \emph{negative}. While at first glance this seems a little odd, there is a good
reason to do so, if one hopes to study systems with more than nearest neighbor coupling.
Suppose for example that one models local interaction as a discretization of a fourth
derivative, a very natural idea. However the couplings to the first and second nearest
neighbors will now have different signs. This goes against the grain of what one
knows about Laplacian systems in general, where in general all couplings must have the same sign
(see \cite{laplacians}). In this case we managed to overcome that problem and analyze stability also when the signs of the (nearest
neighbor) interactions are different. (By necessity they must add up to 1.)
The outline of this paper is as follows. In the next section we start by specifying the model.
Next we discuss the asymptotic stability of the model.
Following \cite{flocks6} we introduce two other types of stability for flocks.
These describe the effect of perturbations in the leaders motion
on the outlying members of the flocks. A flock with $N$ agents is harmonically stable
if the effect of a harmonic motion of the leader on the outlying members
grows less than exponentially fast in $N$ (everything else held fixed).
A flock is said to be impulse stable if the effect of the leader
being kicked is less than exponential on the outlying members. Thus is section
4, we discuss harmonic stability of the model. The problem of impulse stability
is still unsolved. We present a few comments on that problem in section 5.
(The appendix contains technical results and is included for completeness.)
\section{The model} \label{chap:model}
We begin this section establishing the model of this work. The $N+1$ agents move
in $\mathbb{R}$ along orbits $x_i(t)$, $i\in\{0,\cdots n\}$, with velocities
$x_i'(t)$. When they are moving in the desired \emph{formation} their velocities
are equal and their relative positions are determined by $N+1$ a priori given constants $h_i$:
\begin{equation}
x_j-x_{i}= h_j-h_{i} \quad .
\end{equation}
We write the equation of motion for this model in terms of
\begin{equation}
z_i\equiv x_i-h_i \quad .
\end{equation}
These then have the following form:
\begin{equation}\label{example2a}
\ddot z_i = f\left\{z_{i}- (1-\rho)z_{i-1}-\rho z_{i+1}\right\}
+g\left\{\dot z_i-(1-\rho)\dot z_{i-1}-\rho \dot z_{i+1}\right\}\, ,
\end{equation}
for all $i=1,\ldots, N$, and
\begin{equation}\label{example2b}
z_{N+1}(t) = z_0(t)\, ,
\end{equation}
\emph{a priori} given. We will assume the feedback parameters $f$, $g$ are negative reals
and the weight $\rho$ is a arbitrary real number.
It is intuitively convenient, though not necessary, to keep a particular realization
of the above system in mind. Identify $x=N+1$ with $x=0$, so that the agents
move on a (topological) circle. Suppose further that the offsets $h_i$ are given by
$h_i=-i \bmod N$. Now the desired configuration is that of $N+1$ agents moving
at constant speed and uniformly distributed along a circle. (This explains our
title.)
Our strategy here is primarily studying qualitative aspects of the solution of (\ref{example2a})-(\ref{example2b}) as we let $N$ tend to infinity while keeping
all other parameters ($\rho$, $f$, and $g$) fixed.
In particular we wish to understand (1) when the system is asymptotically stable
and (2) how does it converge to its equilibrium when it is asymptotically stable.
This stable equilibrium is given by the
two parameter family of orbits:
$$z_k(t)=z_0(0)+v_0(0)\, t$$ and $$\dot z_k(t)= v_0(0) \quad .$$
These orbits are called \emph{in formation orbits} (for a more detailed discussion, cf. \cite{flocks4, flocks5, flocks6, flocks7}).
It is advantageous to write (\ref{example2a})-(\ref{example2b}) in a more compact form:
$$z\equiv (z_1,\dot z_1, z_2,\dot z_2,\cdots, z_N,\dot z_N)\, .$$
The system can now be recast as a first order ordinary differential equation:
\begin{equation}
\dot z = M z + \Gamma_0(t)\, .
\label{eq:indepleader}
\end{equation}
The matrix $M$ and the vector $\Gamma_0$ are defined below.
Setting
\begin{equation}\label{Q_rho}
Q_\rho=\left(\begin{array}{ccccc}
0 & \rho & & & \\
1-\rho & 0 & \rho & & \\
& \ddots & \ddots & \ddots & \\
& & 1-\rho & 0 & \rho \\
& & & 1-\rho & 0
\end{array}\right)_{N\times N},
\end{equation}
the matrix $P$ defined by
\begin{equation}
P= I-Q_\rho\, ,
\label{eq:laplacian}
\end{equation}
where $I$ is the $N$-dimensional identity matrix, is called the \emph{reduced graph
Laplacian}. It describes the flow of information among the agents, with the exception of the
leader (hence the word `reduced').
The orbit of the leader is assumed to be beforehand given and therefore only appears
in the forcing term $\Gamma_0(t)$.
We will refer to this agent as an \emph{independent leader}.
Analyzing (\ref{example2a})-(\ref{example2b}) and assuming without loss of generality that $h_0=0$, one gathers that:
\begin{equation}
\Gamma_0(t) = \left(\begin{array}{c}
0 \\
(1-\rho)\left(fz_0(t)+g\dot z_0(t)\right) \\
0\\
\vdots \\
0\\
\rho\left(fz_0(t)+g\dot z_0(t)\right)
\end{array}\right) \quad .
\label{eq:Gamma_0}
\end{equation}
In order to define $M$ matrix of (\ref{eq:indepleader}) in terms of these quantities,
we use the Kronecker product, $\otimes$,
$$
M= I \otimes A + P \otimes K \; ,
$$
where $A$ and $K$ the $2\times2$ matrices:
$$
A= \left(\begin{array}{cc}
0 & 1 \\
0 & 0
\end{array}\right) \quad \mbox{ and } \quad
K= \left(\begin{array}{cc}
0 & 0 \\
f & g
\end{array}\right) \quad .
$$
The advantage of this somewhat roundabout way of defining the matrix $M$ is that in the eigenvalues of the reduced Laplacian $P$ can be given explicitly. From that the eigenvalues of $M$ can then be derived.
\section{Asymptotic Stability} \label{chap:asympt}
The system defined in (\ref{example2a})-(\ref{example2b}) is called \emph{asymptotically stable} if all eigenvalues of $M$ have negative real part. Assuming the $\Gamma_0(t)=0$, for $t>t_0$, the solution of the system tends to $0$ exponentially fast (in
$t$) if and only if the system is asymptotically stable. This corresponds to the classical notion of asymptotic stability.
The study of the eigenvalues of the $N\times N$ matrix $Q_\rho$ defined in (\ref{Q_rho})
constitutes a special case of results given in \cite{Fo1,Fo2}. They are given by:
$$2\sqrt{(1-\rho)\rho}\;\cos\left(\frac{\ell\pi}{N+1}\right)\, , \quad \mbox{for}\; \ell=1,2,\ldots,N\, , $$ for all real $\rho$.
These eigenvalues are all real if and only if $\rho\in[0,1]$ and imaginary otherwise, and
the locus of the set of eigenvalues is invariant under multiplication by $-1$. We have:
\begin{proposition}
The reduced Laplacian $P$ has eigenvalues
$\lambda_\ell=1-2\sqrt{(1-\rho)\rho}\, \cos\left(\frac{\ell\pi}{N+1}\right)$,
for $\ell=1,2,\ldots,N$, for all real values of $\rho$.
\end{proposition}
One can show that the eigenvalues of $M=I\otimes A+P\otimes K$ are the solutions $\nu_{\ell\pm}$ of the equation
\begin{equation} \label{eq:evals2}
\nu^2-\lambda_\ell\, g\, \nu-\lambda_\ell\, f= 0 \quad,
\end{equation}
where $\lambda_\ell$ runs through the spectrum of $P$
(cf. \cite{tridiagonal,flocks2,flocks4, flocks5}). So we have:
\begin{theorem} \label{theo:stable}
\begin{enumerate}
\item The eigenvalues of $M$ are
$$
\nu_{\ell\pm} = \frac{1}{2}\left(\lambda_\ell\, g \pm \sqrt{(\lambda_\ell\, g)^2 + 4 \lambda_\ell\, f}\right)\, ,
$$ where $\lambda_\ell$ runs through the spectrum of $P$.
\item For $\rho \in [0,1]$, all real numbers $\lambda_\ell$ are contained in the interval $[0,2]$, and the system is asymptotically stable if and only if both $f$ and $g$ are strictly smaller than zero.
\end{enumerate}
\end{theorem}
These expressions are the same as the corresponding ones for a slghtly differnt
one dimensional flock given in \cite{flocks6}. There it was assumed that $\rho\in[0,1)$.
We extend that research by looking at real values of $\rho$ outside the interval $[0,1]$.
This may at first seem obscure. Here however is the motivation. Suppose for a moment that
one allows each agent to interact with two
neighbors on either side, then one could be tempted to model this interaction as a
discretization of a fourth derivative in the spatial variable. In that case
some of the weights of the interaction would have negative values.
\begin{theorem} \label{theo:stable2}
Let $\rho\in \mathbb{R}\backslash [0,1]$. For a given $f$ and $g$, the system defined in (\ref{example2a})-(\ref{example2b}) is asymptotically
stable, for an arbitrary $N$, if and only if both of the following hold:
\begin{enumerate}
\item $f$ and $g$ are negative, and
\item $f+g^2\geq 0$ or else $4|\rho(1-\rho)|\leq \frac{-g^2}{f+g^2}$.
\end{enumerate}
\end{theorem}
\begin{proof} When $\rho (1-\rho)$ is negative, the eigenvalues $\lambda_\ell$ of $P$
satisfy $\lambda_\ell = 1+i\, a_\ell$, where $a_\ell$ assumes the values
$-2\sqrt{|\rho(1-\rho)|}\;\cos \frac{\ell \pi}{N+1}$.
In particular, for an $N$ sufficiently large, the $a_\ell$'s will distribute themselves smoothly in the interval
\begin{equation}
\left( -2\sqrt{|\rho(1-\rho)|}, +2\sqrt{|\rho(1-\rho)|}\right)\, .
\label{eq:a-interval}
\end{equation}
We need to prove that eigenvalues of $M$ (provided by (\ref{eq:evals2})) have negative
real part. We first look at eigenvalues that correspond to $a_\ell$ approaching to $0$. By
continuity we may set $a_\ell=0$. We get
$$
\nu_{\pm}= \frac{g\pm \sqrt{g^2+4f}}{2} \, .
$$
These roots are real and have opposite signs if $f$ is positive and have the same
sign as $g$ if $f$ is negative. This proves that for $a_\ell$ small enough the corresponding
eigenvalues of $M$ have negative real part if and only if $f$ and $g$ are negative.
\begin{figure}[ptbh]
\centering
\includegraphics[height=3.5in]{flocks8fig1.eps}
\caption{\emph{The calculation of $\nu_\pm$ as function of $a$
for $f=-5$ and $g=-1$, $a$ ranging from 0 to 5.
Notice that $\nu_\pm(0)=0.5(-1\pm i \sqrt{19}\,)$; at $a=0.5$, $\nu_-$ crosses
the imaginary axis. }}
\label{fig:evals}
\end{figure}
It remains to check for what values of $a$ the real part of $\nu_-$ or $\nu_+$ can become
greater than or equal to $0$ (cf. Figure \ref{fig:evals}). To that end we set $\nu=i\tau$, with $\tau$ real. The real and imaginary part of
the equation (\ref{eq:evals2}) now become (abbreviating $a_\ell$ to $a$):
$$
\left\{ \begin{array}{ccc}
\tau^2 - g\, a\, \tau +f &=& 0\\
f\, a + g\,\tau &=& 0 \end{array}
\right.
$$
The first equation defines a hyperbola in the $(a,\tau)$ plane,
and the second equation a line
through the origin. Real solutions for $a$ and $\tau$ exist if and only if $f+g^2<0$ and are given by
$$
(\tau,a)=\frac{\pm 1}{\sqrt{-f-g^2}}\, (f,-g)\, .
$$
Such solutions exist for some $a$ smaller than $2\sqrt{|\rho(1-\rho)|}$
(see interval (\ref{eq:a-interval})) if and only if in addition
$$
2\sqrt{|\rho(1-\rho)|}> \frac{- g}{\sqrt{-f-g^2}}\, .
$$
Finally, if $f+g^2<0$ and $2\sqrt{|\rho(1-\rho)|}>a> \frac{- g}{\sqrt{-f-g^2}}$,
we prove that the system has eigenvalues with positive real part. By continuity,
it is sufficient to prove this only for $a$ arbitrarily large.
In that case Theorem \ref{theo:stable} (1) gives:
\begin{equation}
\nu_-=\frac{\lambda g}{2}\;\left( 1- \sqrt{1+ \frac{4f}{\lambda g^2}} \right)
= -\frac{f}{g} + O(\lambda^{-1}) \quad .
\end{equation}
From equation (\ref{eq:evals2}) we deduce that $\mbox{Re}\, (\nu_+)+\mbox{Re}\, (\nu_-)=\mbox{Re}\, (g(1+ia))=g$.
Thus as $a$ tends to infinity, $\mbox{Re}\, (\nu_+f)=g+\frac fg >0$, which was to be proved.
\end{proof}
\section{Harmonic Stability} \label{chap:harmonic}
The system is harmonically unstable roughly if oscillatory or harmonic perturbations in the orbit
of the leader (that is: of the form $e^{i\omega t}$) have their amplitude magnified by a
factor that is exponentially large in $N$ (cf. \cite{flocks6}).
We first need some notation. It will often be convenient to replace
$\rho$ by a different constant:
$$
\kappa = \frac{1-\rho}{\rho}
$$
or, equivalently,
$$\rho = \frac{1}{1+\kappa} \, ,$$
We also define (for $\rho \neq 0$):
\begin{equation}\label{mu_pm}
\mu_\pm \equiv \frac{1}{2\rho}\left( \gamma\pm \sqrt{\gamma^2-4\rho(1-\rho)}\right)
\end{equation}
where
$$
\gamma = \frac{f+i\,\omega\, g +\omega ^2}{f+i\,\omega\, g}\, .$$
\begin{proposition}\label{prop:a_n}
The frequency response function of the $k$-th agent is given by
$$
a_{k}(f,g,\omega)=\frac{\kappa^k(\mu_-^{N+1-k}-\mu_+^{N-k})+(\mu_-^{k}-\mu_+^{k})} {\mu_+^{N}-\mu_-^{N}}\, ,
$$
where $\mu_\pm$ is defined as in (\ref{mu_pm}).
\end{proposition}
\begin{proof} All eigenvalues of $M$ have negative real part. Let $z_0(t)$ be given by $e^{i\omega t}$. Under these assumptions, the motion of the system is asymptotic (as $t\rightarrow\infty$) to $z_k=a_k\,e^{i\omega t}$. This leads to a recursive equation on $a_k$, for $k=1,\ldots, N$,
$$
(1-\rho)\, a_{k-1}-\gamma \, a_k +\rho \, a_{k+1}=0 \, .
$$
The boundary conditions are given by:
$$
a_0 = a_{N+1}=1 \, .
$$
Let $\mu_\pm$ be the roots of the associated characteristic polynomial
$$P(x)= x^2-\frac\gamma\rho\, x +\frac{1-\rho}{\rho}\, .$$
The general solution is $$a_k=c_-\mu_-^k+c_+\mu_+^k\, .$$ A convenient way to solve for $c_\pm$ is by setting $d_1=c_-\mu_-^N$ and $d_2=c_+\mu_+^N$. The boundary conditions can be rewritten as
$$
\left( \begin{array}{cc} 1& 1\\
\mu_-^{N+1}& \mu_+^{N+1}\end{array} \right)
\left( \begin{array}{c} c_-\\ c_+ \end{array} \right) =
\left( \begin{array}{c} 1\\ 1 \end{array} \right)$$ or, equivalently,
$$
\left( \begin{array}{c} c_-\\ c_+ \end{array} \right) =
\frac{1}{\mu_+^{N+1}-\mu_-^{N+1}} \;
\left( \begin{array}{c} \mu_+^{N+1}-1 \\1-\mu_-^{N+1} \end{array} \right)\, .
$$
Substituting this into $a_k$ and using the fact that the product of the $\mu_\pm$ equals $\kappa$, we get the result.
\end{proof}
We will assume here without further proof that fluctuations of the leader that
are propagated through the system are largest for the agents furthest away from the
leader, \emph{i.e.} halfway in the flock. For simplicity we only consider in this section
the response of the agent $k=\frac{N+1}{2}$ where $N$ is odd (and large).
\begin{corollary} \label{cory:halfway}
If $N$ is odd, we have for $M=\frac{N+1}{2}$:
$$
a_{M}(f,g,\omega)=\frac{(\kappa^M+1)} {\mu_-^{M}+\mu_+^{M}}\, ,
$$
where $\mu_\pm$ is defined as in (\ref{mu_pm}).
\end{corollary}
Recall that by Theorems \ref{theo:stable} and \ref{theo:stable2} the system is asymptotically
stable if and only if $\rho\in[0,1]$ and $f$ and $g$ negative \emph{or else} $\rho\in \mathbb{R}\backslash [0,1]$ and both of the following hold:
\begin{itemize}
\item $f$ and $g$ are negative, and
\item $f+g^2\geq 0$ or else $4|\rho(1-\rho)|\leq \frac{-g^2}{f+g^2}$.
\end{itemize}
Recall from the proof of Theorem \ref{prop:a_n} that if $z_0(t)$ equals $e^{i\omega t}$ then $z_k$ is asymptotic to $a_k\,e^{i\omega t}$.
Thus the amplification at the $k$-th agent of the leader's signal is
given $a_k(\omega)$. We need to determine whether $\max_k \sup_{\omega}|a_k(\omega))|$ is exponential
in $N$ (instability) or less than exponential (stability). We may assume $k=M$. So let $$A_M\equiv \sup_{\omega\in\mathbb{R}}\; |a_M(i\omega)|\, .$$ Following \cite{flocks6}, we call a system \emph{harmonically stable} if it is asymptotically stable and if $$\limsup_{M\rightarrow \infty}\; \left|A_M\right|^{1/N}\leq 1\, .$$
\begin{theorem}
The system given by the equation (\ref{eq:indepleader}) is harmonically stable if and only if
$\rho=1/2$ and $f$ and $g$ are negative.
\label{theo:harmonic}
\end{theorem}
\begin{figure}[ptbh]
\centering
\includegraphics[height=2.3in]{flocks8fig3.eps}
\caption{\emph{ The eigenvalues $\mu_+(\omega)$ (blue) and $\mu_-(\omega)$ (red)
when $f=g=-1$ and $\rho=1/2$, for $\omega$ positive. In addition the unit circle is drawn in green.}}
\label{fig:rho=1/2}
\end{figure}
\begin{proof}
We concentrate first on the $\rho=1/2$ case. Here we have:
$$
a_M(\omega) = \frac{2}{\mu_+(\omega)^M + \mu_-(\omega)^M} \quad \mbox{ and } \quad \mu_+\mu_-=1 \, .
$$
Geometrically what happens is that the system exhibits near-resonance. More precisely,
the curves $\mu_\pm(\omega)$ are (quadratically) tangent to the unit circle at $\omega=0$
(see Figure \ref{fig:rho=1/2}). Of course when $\mu_\pm(\omega)=e^{\pm i\pi/2M}$ the denominator
cancels and $a_M$ is undefined. The quadratic tangency means that (for $M$ large) the curves
$\mu_\pm(\omega)$ pass the points $e^{\pm i\pi/2M}$ on the unit circle at a distance proportional
to $1/M$. In turn this means that
$$
A_M = \sup_{\omega}\, |a_M(\omega))|
$$
grows \emph{linearly} in $M$ and is thus harmonically stable.
Analytically this can be worked out precisely by doing a pole expansion on $a_M(\omega)$.
A very similar calculation was done in detail in \cite{flocks5} and we will not repeat
that calculation here. The only differences with that calculation are:
here we are calculating $a_M$ and not $a_N$, and here our eigenvalues $\nu_{\ell\pm}$ are slightly different from those in the cited paper.
Now we turn to the other cases: $\rho\neq 1/2$. We first argue that the cases $\rho$ and $1-\rho$
are symmetric. In particular if we use a new value $\rho'=1-\rho$ instead of $\rho$, then
in the expression for $a_M$, $\mu_\pm$ and $\kappa$ are all replaced by their reciprocals
as can be seen by inspecting the polynomial $P$ in the proof of Proposition \ref{prop:a_n}.
A little calculation shows that $a_M$ is invariant under this operation. It is thus sufficient
to consider only $\rho<1/2$.
Proposition \ref{prop:mu+bigger} tells us that for all $\rho<1/2$ there is no resonance
or near resonance as the two $\mu$ have distinct modulus. Lemma \ref{lem:omega+} implies that,
for $\omega$ less than some $\omega+$ given there, $\mu_-(\omega)$ is bigger than 1. Since
in this case $|\kappa|>1$ and $\mu_+\mu_-=\kappa$, we have for $\omega \in (0,\omega+)$:
$$
|\kappa|>|\mu_+(\omega)|>|\mu_-(\omega)| \quad ,
$$
and thus the expression in the above Corollary grows exponentially in $M$. Therefore all
these systems are harmonically unstable.
\end{proof}
\section{An open problem} \label{chap:impulse}
Suppose now that the flock is in a stable equilibrium (\emph{i.e.} moves stably in formation) when the
leader suddenly and quickly changes its velocity. Roughly speaking we call the system
\emph{impulse stable} when the physical response of the other agents (\emph{i.e.},
the acceleration, or the velocity, or the position) is less than exponential in $M$.
As observed in the proof of Theorem \ref{theo:harmonic}, the case $\rho=\frac12$
is extremely similar to the problem studied in \cite{flocks5}, and the solution is in fact
similar to the one given in that case. The calculations there indicate responses that
are `proportional' to $M$. Thus may we conclude that here also:
\begin{proposition} The system given in the equation (\ref{eq:indepleader}) is impulse
stable when it is asymptotically stable and when $\rho=1/2$.
\end{proposition}
The cases $\rho\neq 1/2$ lead to problems similar to the one that remained unsolved
in \cite{flocks7}. It seems very likely at this point that most of these cases are
impulse unstable, though this is by no means obvious or known. More specifically,
as we vary $f$, $g$ and $\rho\neq 1/2$, we do not even qualitatively understand the large $N$ behavior of the motion of these flocks. This question is relevant because
velocity changes of the leader are a natural context in which stability plays
an important role for the cohesion of the flock. We might think for example of a lead car
accelerating when a traffic light turns green or a large flock of animals changing course
because outlying members spotted and try evade a predator.
The mathematical problem boils down to an inverse Fourier transform of $a_M(\omega)$
where $M$ is large. Current
standard integration techniques do not readily give asymptotic
(in $M$) expressions for such integrals. We do not address this challenging question any further here, leaving it as an open problem for future research.
\section{APPENDIX: Technical Results} \label{chap:tresults}
In this section we gather some technical results which we exhibited partially before
in \cite{flocks6}. The results there were proved only for $\rho\in[0,1]$.
Some of the calculations extend verbatim (or almost) to all $\rho\in\mathbb{R}$. The proof
of the main result, Proposition \ref{prop:mu+bigger}, had to be modified substantially however.
\begin{lemma} \label{lem:taylor} Let $\omega \geq 0$ be sufficiently small.
\begin{enumerate}
\item For $\rho\in (0,\frac12)$, we have:
$$
\mu_+ = \frac{1-\rho}{\rho}\left(1 + \frac{\omega^2}{(2\rho-1)|f|} - i\, \frac{|g| \,\omega^3}{(2\rho-1)f^2} \right) + \mathcal{O}(\omega^4)$$ and
$$ \mu_- = 1 - \frac{\omega^2}{(2\rho-1)|f|} + i\, \frac{|g| \, \omega^3}{(2\rho-1)f^2} + \mathcal{O}(\omega^4) \, .$$
\item For $\rho\in (\frac12,1)$, we have
$$
\mu_+ = 1 - \frac{\omega^2}{(2\rho-1)|f|} + i\; \frac{|g| \, \omega^3}{(2\rho-1)f^2} + \mathcal{O}(\omega^4) $$ and
$$
\mu_- = \frac{1-\rho}{\rho}\left(1 + \frac{\omega^2}{(2\rho-1)|f|} - i\, \frac{|g| \, \omega^3}{(2\rho-1)f^2} \right) + \mathcal{O}(\omega^4) \, .
$$
\end{enumerate}
\end{lemma}
\begin{proof} By sheer calculation. (See \cite{flocks4} for some of the computational details.) \end{proof}
\begin{remark} This expansion diverges for $\rho=1/2$; in that case we have (cf. \cite{flocks4}):
$$
\mu_\pm = 1-\frac{\omega^2}{|f|}\pm
\frac{\omega^2|g|}{\sqrt{2}|f|^{3/2}}+\mathcal{O}(\omega^4) +i\left( \pm
\frac{\sqrt{2}\omega}{|f|^{1/2}} \pm \mathcal{O}(\omega^3) \right)
\, .
$$
\label{remark}
\end{remark}
\begin{lemma}\label{lem:gamma} $$\gamma(\omega)=1-\frac{\omega^2 |f|}{f^2+\omega^2g^2} + i\,\frac{\omega^3|g|}{f^2+\omega^2g^2}\, .$$
\end{lemma}
The complicated looking conditions in the following proposition are
nothing but the conditions that insure asymptotic stability (see Theorems
\ref{theo:stable} and \ref{theo:stable2}).
\begin{proposition} \label{prop:mu+bigger}
Let $\rho\in[0,1]$ and $f$ and $g$ negative or let $\rho\in \mathbb{R}\backslash [0,1]$
and suppose that both of the following hold:
\begin{description}
\item[i] $f$ and $g$ are negative, and
\item[ii] $f+g^2\geq 0$ or else $4|\rho(1-\rho)|\leq \frac{-g^2}{f+g^2}$.
\end{description}
Then, $r$ defined by $$r=\sup_{\omega>0} \frac{|\mu_-(\omega)|}{|\mu_+(\omega)|}$$ exists and is contained in interval $(0,1)$ with only two exceptions:
\begin{itemize}
\item when $\rho=\frac12$, $\frac{|\mu_-(\omega)|}{|\mu_+(\omega)|}$ equals $1$ at $\omega =0$ and is
is strictly smaller than $1$ for $\omega>0$, or
\item when $4\rho(1-\rho)= \frac{-g^2}{f+g^2}<0<0$, $\frac{|\mu_-(\omega)|}{|\mu_+(\omega)|}$ is strictly
smaller than $1$, if $0\geq\omega^2<\frac{f^2}{-f-g^2}$, and $|\mu_-(\omega)|=|\mu_+(\omega)|$, if
$\omega^2=\frac{f^2}{-f-g^2}$.
\end{itemize}
\end{proposition}
\begin{figure}[ptbh]
\centering
\includegraphics[height=2.3in]{flocks8fig2.eps}
\caption{\emph{ The constant $\kappa=(1-\rho)/\rho$ as function of $\rho$.}}
\label{fig:rho}
\end{figure}
\begin{proof} It is clear from remark \ref{remark} that when $\rho=\frac12$ the two eigenvalues are equal to $1$,
when $\omega=0$, and so at that point the quotient equals $1$.
The arguments below establish that the quotient is always smaller than $1$, for all positive values of $\omega$.
As far as the second exception is concerned: We will show that the equality of $\mu_+$ and
$\mu_-$ occurs at some $\omega>0$. The arguments below, however, insure that for smaller (non-negative)
values of $\omega$, the modulus of the quotient of the eigenvalues is strictly smaller than $1$.
From its definition, $\gamma(\omega)\approx \frac{\omega}{i g}$ when $\omega$ is large. Substitute this into the expression for $\mu_\pm$ in equation (\ref{mu_pm}) to see that for a large enough $\omega$, in fact, $\frac{|\mu_-(\omega)|}{|\mu_+(\omega)|}$ becomes very small.
In the following, note that $|\kappa|>1$ if and only if $\rho<\frac12$ and
$|\kappa|>1$ if and only if $\rho>\frac12$ (see Figure \ref{fig:rho}).
So when $\omega=0$, Lemma \ref{lem:taylor} implies that $\frac{|\mu_-|}{|\mu_+|}$ equals $\min\{|\kappa|,|\kappa|^{-1}\}$, for all $\rho$.
It is now sufficient to prove that, for $\omega\in\mathbb{R}^+$, the absolute values $|\mu_\pm|$ are never equal. So suppose there are $\omega_0$ and $\theta\in\mathbb{R}$ such that $\mu_+(\omega_0)-\mu_-(\omega_0)e^{i\theta}=0$. The definition of $\mu_\pm$ in equation (\ref{mu_pm}) provides:
$$
\gamma (1-e^{i\theta}) = -\sqrt{\gamma^2-4\rho(1-\rho)}\,(1+e^{i\theta})\, .
$$
Dividing this by $1+e^{i\theta}$, squaring the equation, and noting that $$\frac{(1-e^{i\theta})^2}{(1+e^{i\theta})^2}=-\tan^2\left(\frac\theta2\right)\, ,$$ we get
$$
\gamma^2 \left(1+\tan^2 \frac{\theta}{2}\right)=4\rho(1-\rho) \, .
$$
If $\rho(1-\rho)>0$, then $\gamma^2$ is a positive real number and therefore $\gamma$ is real for some $\omega\neq 0$, which is impossible by Lemma \ref{lem:gamma}. If $\rho(1-\rho)<0$, then
$\gamma^2$ is a negative real number so that $\gamma$ is imaginary. Setting the real part of $\gamma$ equal
to $0$ in Lemma \ref{lem:gamma} yields $f^2+\omega^2(f+g^2)=0$. If $f+g^2$ is non-negative, this
has no solution (because $\omega$ is real). So suppose it is positive. Then substitute the positive
solution into $\gamma$ and check that $\gamma=i\frac{|g|}{\sqrt{-f-g^2}}$.
Substituting this in turn into (\ref{mu_pm}), we see that the modulus of $\mu_+$ is greater
than that of $\mu_-$ (here $\sqrt{\cdot}$ means the root in the \emph{upper} half plane)
as long as $4|\rho(1-\rho)|< \frac{-g^2}{f+g^2}$. When $4|\rho(1-\rho)|= \frac{-g^2}{f+g^2}$, we have $\mu_+=\mu_-$.
\end{proof}
\begin{figure}[ptbh]
\centering
\includegraphics[height=2.3in]{flocks8fig4.eps}
\includegraphics[height=2.3in]{flocks8fig5.eps}
\caption{\emph{An illustration of the curious behavior of $\mu_\pm(\omega)$.
The eigenvalues $\mu_+(\omega)$ are in blue and $\mu_-(\omega)$ are in red,
when $f=-5$ and $g=-1$, for positive $\omega$. The circle $r=\kappa$ (green),
$r=\sqrt{|\kappa|}$ (yellow), and the unit circle (black) are also drawn.
Note that $\mu_+(0)=\kappa$ (which is negative in both cases) and that $\mu_-(0)=1$.
In the first figure $\rho=-0.05$ and $\mu_{blue}$ is always bigger than $\mu_{red}$ except
MAPLE insisted in using the principal root for the square root as opposed to our convention,
and so it recklessly swaps the $2$ roots. In the second picture $4\rho(1-\rho)> \frac{-g^2}{f+g^2}<0$
(while $\rho(1-\rho) < 0$) and now the quotient of the two roots crosses $1$. }}
\label{fig:rho-extra}
\end{figure}
\begin{lemma} \label{lem:omega+} For each $\rho < 1/2$, there is a unique $\omega_+>0$ such that
$$
\omega\in\left(0,\omega_+\right)\quad \mbox{ implies } \quad |\mu_-(\omega)|>1$$
and
$$
\omega>\omega_+\quad \mbox{ implies }\quad|\mu_-(\omega)|<1 \,.
$$
\end{lemma}
\begin{proof} We know that $\mu_-(0)=1$ and, from the proof of the previous lemma, for a large $\omega$, $|\mu_-(\omega)|$ is small. It is sufficient to prove that $\omega_+$ is the unique solution in $(0,\infty)$ of $|\mu_-(\omega)|=1$ and that it is simple.
In fact, consider the characteristic equation
$\rho\, \mu^2-\gamma\, \mu + (1-\rho)=0$ and suppose that there is a root $\mu=e^{i\theta}$. Then $$\gamma=\rho e^{i\theta}+(1-\rho)e^{-i\theta} =\cos(\theta)+i(2\rho-1)\sin(\theta)\, .$$ Equating this to the expression given in Lemma \ref{lem:gamma} and using the Pythagorean trigonometric identity, we to obtain:
$$
\left(1-\frac{\omega^2 |f|} {f^2+\omega^2g^2}\right)^2+\frac{1}{(2\rho-1)^2}\left(\frac{\omega^3|g|} {f^2+\omega^2g^2}\right)^2=1
$$
This equation factors as follows:
$$
\omega^2\left(\frac{g^2}{(2\rho-1)^2}\,\omega^4+(f^2-2|f|g^2)\omega^2-2|f|^3 \right)=0
$$
The second factor is a quadratic expression in $\omega^2$ which has a positive leading
coefficient and a negative trailing coefficient. This gives exactly one simple positive root for $\omega^2$, yielding a unique simple positive root $\omega=\omega_+$. \end{proof}
\begin{remark} In fact,
$$
\omega_+^2= (1-2\rho)^2|f|\left( 1-\frac{|f|}{2g^2}
+ \sqrt{\left(1-\frac{|f|}{2g^2}\right)^2 +\,\frac{2|f|}{(1-2\rho)^2g^2}} \right) \, .
\label{eq:omega+}
$$
\end{remark}
|
\section{An Introduction}
Quantum systems with variable quadratic Hamiltonians are called the
generalized harmonic oscillators (see, for example, \cite{Berry85}, \cit
{Dod:Mal:Man75}, \cite{Dodonov:Man'koFIAN87}, \cite{Hannay85}, \cite{Leach90
, \cite{Malkin:Man'ko79}, \cite{Wolf81}, \cite{Yeon:Lee:Um:George:Pandey93}
and references therein). They have attracted substantial attention over the
years in view of their great importance in many advanced quantum problems.
Examples include coherent states and uncertainty relations \cit
{Malkin:Man'ko79}, \cite{Malk:Man:Trif69}, \cite{Malk:Man:Trif70}, \cit
{Mand:Karpov:Cerf}, \cite{Klauder:Sudarshan}, Berry's phase \cite{Berry85},
\cite{Berry:Hannay88}, \cite{Hannay85}, \cite{Leach90}, \cite{Morales88},
asymptotic and numerical methods \cite{Kruskal62}, \cite{Maslov:Fedoriuk},
\cite{Milne30}, \cite{Mun:Ru-Paz:Wolf09}, quantization of mechanical systems
\cite{Faddeyev69}, \cite{FeynmanPhD}, \cite{Feynman}, \cite{Fey:Hib}, \cit
{Kochan07}, \cite{Kochan10} and Hamiltonian cosmology \cit
{Bertoni:Finelli:Venturi98}, \cite{Finelli:Gruppuso:Venturi99}, \cit
{Finelli:Vacca:Venturi98}, \cite{Hawkins:Lidsey02}, charged particle traps
\cite{Major:Gheorghe:Werth} and motion in uniform magnetic fields \cit
{Cor-Sot:Lop:Sua:Sus}, \cite{Corant:Snyder58}, \cite{Dodonov:Man'koFIAN87},
\cite{La:Lif}, \cite{Lewis67}, \cite{Lewis68}, \cite{Lewis:Riesen69}, \cit
{Malk:Man:Trif70}, molecular spectroscopy \cite{Dokt:Mal:Man77}, \cit
{Malkin:Man'ko79} and polyatomic molecules in varying external fields,
crystals through which an electron is passing and exciting the oscillator
modes, and other interactions of the modes with external fields \cit
{Fey:Hib}. Quadratic Hamiltonians have particular applications in quantum
electrodynamics because the electromagnetic field can be represented as a
set of forced harmonic oscillators \cite{Bo:Shi}, \cite{Dodonov:Man'koFIAN87
, \cite{Fey:Hib}, \cite{Merz} and \cite{Schiff}. Nonlinear oscillators play
a central role in the novel theory of Bose--Einstein condensation \cit
{Dal:Giorg:Pitaevski:Str99}. From a general point of view, the dynamics of
gases of cooled atoms in a magnetic trap at very low temperatures can be
described by an effective equation for the condensate wave function known as
the Gross--Pitaevskii (or nonlinear Schr\"{o}dinger) equation \cit
{Kagan:Surkov:Shlyap96}, \cite{Kagan:Surkov:Shlyap97}, \cit
{Kivsh:Alex:Tur01} and \cite{Per-G:Tor:Mont}.\medskip\
In this Letter we consider the one-dimensional time-dependent Schr\"{o
dinger equatio
\begin{equation}
i\frac{\partial \psi }{\partial t}=H\psi \label{Schroedinger}
\end{equation
with general variable quadratic Hamiltonians of the for
\begin{equation}
H=a\left( t\right) p^{2}+b\left( t\right) x^{2}+c\left( t\right) px+d\left(
t\right) xp,\quad p=-i\frac{\partial }{\partial x}, \label{GenHam}
\end{equation
where $a\left( t\right) ,$ $b\left( t\right) ,$ $c\left( t\right) ,$ and
d\left( t\right) $ are real-valued functions of time, $t,$ only (see, for
example, \cite{Cor-Sot:Lop:Sua:Sus}, \cite{Cor-Sot:Sua:Sus}, \cit
{Cor-Sot:Sua:SusInv}, \cite{Cor-Sot:Sus}, \cite{Dod:Mal:Man75}, \cit
{FeynmanPhD}, \cite{Feynman}, \cite{Fey:Hib}, \cite{Lan:Sus}, \cite{Lop:Sus
, \cite{Me:Co:Su}, \cite{SuazoF}, \cite{Sua:Sus}, \cite{Suaz:Sus}, \cit
{Sua:Sus:Vega}, \cite{Wolf81}, and \cite{Yeon:Lee:Um:George:Pandey93} for a
general approach and some elementary solutions). The corresponding Green
function, or Feynman's propagator, can be found as follows \cit
{Cor-Sot:Lop:Sua:Sus}, \cite{Suaz:Sus}
\begin{equation}
\psi =G\left( x,y,t\right) =\frac{1}{\sqrt{2\pi i\mu _{0}\left( t\right) }}\
e^{i\left( \alpha _{0}\left( t\right) x^{2}+\beta _{0}\left( t\right)
xy+\gamma _{0}\left( t\right) y^{2}\right) }, \label{in2}
\end{equation
wher
\begin{eqnarray}
&&\alpha _{0}\left( t\right) =\frac{1}{4a\left( t\right) }\frac{\mu
_{0}^{\prime }\left( t\right) }{\mu _{0}\left( t\right) }-\frac{c\left(
t\right) }{2a\left( t\right) }, \label{in3} \\
&&\beta _{0}\left( t\right) =-\frac{\lambda \left( t\right) }{\mu _{0}\left(
t\right) },\qquad \lambda \left( t\right) =\exp \left( \int_{0}^{t}\left(
c\left( s\right) -d\left( s\right) \right) \ ds\right) , \label{in4} \\
&&\gamma _{0}\left( t\right) =\frac{a\left( t\right) \lambda ^{2}\left(
t\right) }{\mu _{0}\left( t\right) \mu _{0}^{\prime }\left( t\right) }+\frac
c\left( 0\right) }{2a\left( 0\right) }-4\int_{0}^{t}\frac{a\left( s\right)
\sigma \left( s\right) \lambda ^{2}\left( s\right) }{\left( \mu _{0}^{\prime
}\left( s\right) \right) ^{2}}\ ds, \label{in5}
\end{eqnarray
and the function $\mu _{0}\left( t\right) $ satisfies the\ characteristic
equatio
\begin{equation}
\mu ^{\prime \prime }-\tau \left( t\right) \mu ^{\prime }+4\sigma \left(
t\right) \mu =0 \label{in6}
\end{equation
wit
\begin{equation}
\tau \left( t\right) =\frac{a^{\prime }}{a}+2c-2d,\qquad \sigma \left(
t\right) =ab-cd+\frac{c}{2}\left( \frac{a^{\prime }}{a}-\frac{c^{\prime }}{c
\right) \label{in7}
\end{equation
subject to the initial dat
\begin{equation}
\mu _{0}\left( 0\right) =0,\qquad \mu _{0}^{\prime }\left( 0\right)
=2a\left( 0\right) \neq 0. \label{in8}
\end{equation
(More details can be found in Refs.~\cite{Cor-Sot:Lop:Sua:Sus} and \cit
{Suaz:Sus}.) Then by the superposition principle the solution of the Cauchy
initial value problem can be presented in an integral for
\begin{equation}
\psi \left( x,t\right) =\int_{-\infty }^{\infty }G\left( x,y,t\right) \
\varphi \left( y\right) \ dy,\quad \lim_{t\rightarrow 0^{+}}\psi \left(
x,t\right) =\varphi \left( x\right) \label{CauchyInVProb}
\end{equation
for a suitable initial function $\varphi $ on
\mathbb{R}
$ (a rigorous proof is given in \cite{Suaz:Sus} and uniqueness is analyzed
in \cite{Cor-Sot:Sua:SusInv}; another forms of solution are provided by (\re
{KSolution}) and (\ref{QIExpSolution})).\medskip
A detailed review on dynamical symmetries and quantum integrals of motion
for the time-dependent Schr\"{o}dinger equation can be found in \cit
{Dodonov:Man'koFIAN87} and \cite{Malkin:Man'ko79} (see also an extensive
list of references therein). In this Letter, which is a continuation of the
recent paper \cite{Cor-Sot:Sua:SusInv}, a natural connection between the
linear and quadratic integrals of motion for general variable quadratic
Hamiltonians is established. As a result a nonlinear superposition principle
for the corresponding Ermakov systems, known as Pinney's solution, is
obtained and generalized. We pay also special attention to fundamental
relations between the linear dynamical invariants of Dodonov, Malkin, Man'ko
and Trifonov and solutions of the Cauchy initial value problem (see original
works \cite{Dod:Mal:Man75}, \cite{Dodonov:Man'koFIAN87}, \cit
{Malkin:Man'ko79}, \cite{Malk:Man:Trif73}). In addition this initial value
problem is explicitly solved in terms of the quadratic invariant
eigenfunction expansion, which seems to be missing in the available
literature in general.\medskip
The Letter is organized as follows. We start from the standard definitions
of the dynamical symmetry and quantum integrals of motion, introducing also
some elementary tools, in the next two sections. Then in Section~4 we
describe all linear dynamical invariants for variable quadratic Hamiltonians
and determine their actions on the solutions of the corresponding
time-dependent Schr\"{o}dinger equation. In Section~5 all quadratic quantum
integrals of motion are characterized in terms of solutions of the
generalized Ermakov equation and a detailed proof is given. Their
decomposition into products of the linear invariants is derived in Section~6
and explicit actions on the solutions are established in Section~7. The last
section deals with a related nonlinear superposition principle for Ermakov's
equations and some computational details are provided in Appendices~A and B.
\section{Dynamical Symmetry}
In this Letter we elaborate on the following property.
\begin{lemma}
I
\begin{equation}
i\frac{\partial \psi }{\partial t}=H\psi ,\qquad i\frac{\partial O}{\partial
t}+OH-H^{\dagger }O=0, \label{Invariant}
\end{equation
then the function $\psi _{1}=O\psi $ satisfies the time-dependent Schr\"{o
dinger equatio
\begin{equation}
i\frac{\partial \psi _{1}}{\partial t}=H^{\dagger }\psi _{1},
\label{AdjointSchroedinger}
\end{equation
where $H^{\dagger }$ is the adjoint of the Hamiltonian $H.$
\end{lemma}
When $H=H^{\dagger },$ this property is taken as a definition of the
dynamical symmetry of the time-dependent Schr\"{o}dinger equation (\re
{Schroedinger}) (see, for example, \cite{Dodonov:Man'koFIAN87}, \cit
{Lewis:Riesen69}, \cite{Malkin:Man'ko79} and references therein). At the
same time one has to deal with non-self-adjoint Hamiltonians in the theory
of dissipative quantum systems (see, for example, \cite{Cor-Sot:Sua:Sus},
\cite{Dekker81}, \cite{Kochan10}, \cite{Tarasov01}, \cite{Um:Yeon:George}
and references therein) or when using separation of variables in an
accelerating frame of reference for a charged particle moving in an uniform
variable magnetic field \cite{Cor-Sot:Lop:Sua:Sus}.
\begin{proof}
Partial differentiatio
\begin{eqnarray}
&&i\frac{\partial \psi _{1}}{\partial t}=i\frac{\partial }{\partial t}\left(
O\psi \right) =i\frac{\partial O}{\partial t}\psi +iO\frac{\partial \psi }
\partial t} \\
&&\qquad \quad =\left( H^{\dagger }O-OH\right) \psi +OH\psi =H^{\dagger
}\psi _{1} \notag
\end{eqnarray
provides a direct proof.
\end{proof}
Definition of the dynamical symmetry is usually given in terms of solutions
of the same equation. A simple modification helps with the non-self-adjoint
quadratic Hamiltonians.
\begin{lemma}
The wave functions $\psi $ and $\chi ,$ related b
\begin{equation}
\psi =\left( e^{-\int_{0}^{t}\left( c-d\right) ds}\ O\right) \chi ,
\label{Conjugate}
\end{equation
are solutions of the same Schr\"{o}dinger equation (\ref{Schroedinger})--
\ref{GenHam}), if the operator $O$ satisfies hypothesis of Lemma~1.
\end{lemma}
\begin{proof}
The simplest dynamical invariant, or an operator with the property (\re
{Invariant}), is given b
\begin{equation}
O_{0}=O_{0}\left( c,d\right) =e^{\int_{0}^{t}\left( c-d\right) ds}\ I,
\label{SimplestInvariant}
\end{equation
where $I=id$ is the identity operator. (More details are provided in
Section~4.) Apply Lemma~1 twice in the following orde
\begin{equation}
\psi =O_{0}\left( d,c\right) \left( O\chi \right)
\end{equation
and use $\left( H^{\dagger }\right) ^{\dagger }=H$ to complete the proof.
\end{proof}
Examples occur throughout the Letter.
\section{Differentiation of Operators and Dynamical Invariants}
Following Lemma~1 we define the time derivative of an operator $O$ as follow
\begin{equation}
\frac{dO}{dt}=\frac{\partial O}{\partial t}+\frac{1}{i}\left( OH-H^{\dagger
}O\right) , \label{Diff}
\end{equation
where $H^{\dagger }$ is the adjoint of the Hamiltonian operator $H.$ (This
formula is a simple extension of the well-known expression \cite{Deb:Mikus},
\cite{La:Lif}, \cite{Merz}, \cite{Schiff} to the case of a non-self-adjoint
Hamiltonian \cite{Cor-Sot:Sua:Sus}.) By definitio
\begin{equation}
\frac{dO}{dt}=\frac{\partial O}{\partial t}+\frac{1}{i}\left( OH-H^{\dagger
}O\right) =0 \label{DynamicalInvariant}
\end{equation
for any dynamical invariant. (In what follows we assume that the invariant
O $ does not involve time differentiation.)\medskip
This derivative is a linear operato
\begin{equation}
\frac{d}{dt}\left( c_{1}O_{1}+c_{2}O_{2}\right) =c_{1}\frac{dO_{1}}{dt}+c_{2
\frac{dO_{2}}{dt} \label{DiffLinear}
\end{equation
and the product rule takes the for
\begin{eqnarray}
\frac{d}{dt}\left( O_{1}O_{2}\right) &=&\frac{\partial \left(
O_{1}O_{2}\right) }{\partial t}+\frac{1}{i}\left( \left( O_{1}O_{2}\right)
H-H^{\dagger }\left( O_{1}O_{2}\right) \right) \label{ProductRule} \\
&=&\frac{dO_{1}}{dt}O_{2}+O_{1}\frac{dO_{2}}{dt}+iO_{1}\left( H-H^{\dag
}\right) O_{2}. \notag
\end{eqnarray
For the general quadratic Hamiltonian, (\ref{GenHam}), one get
\begin{equation}
\frac{d}{dt}\left( O_{1}O_{2}\right) =\frac{dO_{1}}{dt}O_{2}+O_{1}\frac
dO_{2}}{dt}+\left( c-d\right) O_{1}O_{2} \label{ProductRuleA}
\end{equation
and by definition (\ref{Diff}
\begin{eqnarray}
\frac{d}{dt}\left( e^{\alpha \int_{0}^{t}\left( c-d\right) ds}\
O_{1}O_{2}\right) &=&e^{\alpha \int_{0}^{t}\left( c-d\right) ds}\ \left(
\frac{dO_{1}}{dt}O_{2}+O_{1}\frac{dO_{2}}{dt}\right) \\
&&+\left( \alpha +1\right) \left( c-d\right) e^{\alpha \int_{0}^{t}\left(
c-d\right) ds}\ O_{1}O_{2}. \notag
\end{eqnarray
If $\alpha =-1,$ we finally obtai
\begin{equation}
\frac{d}{dt}\left( e^{-\int_{0}^{t}\left( c-d\right) ds}\ O_{1}O_{2}\right)
=e^{-\int_{0}^{t}\left( c-d\right) ds}\ \left( \frac{dO_{1}}{dt}O_{2}+O_{1
\frac{dO_{2}}{dt}\right) . \label{productrule}
\end{equation
This implies that, if the operators $O_{1}$ and $O_{2}$ are dynamical
invariants, namely
\begin{eqnarray}
&&\frac{dO_{1}}{dt}=\frac{\partial O_{1}}{\partial t}+\frac{1}{i}\left(
O_{1}H-H^{\dagger }O_{1}\right) =0, \\
&&\frac{dO_{2}}{dt}=\frac{\partial O_{2}}{\partial t}+\frac{1}{i}\left(
O_{2}H-H^{\dagger }O_{2}\right) =0,
\end{eqnarray
then their modified product
\begin{equation}
E=e^{-\int_{0}^{t}\left( c-d\right) ds}\ O_{1}O_{2},
\label{dynamicinvariant}
\end{equation
is also a dynamical invariant
\begin{equation}
\frac{dE}{dt}=\frac{\partial E}{\partial t}+\frac{1}{i}\left( EH-H^{\dagger
}E\right) =0.
\end{equation
In Section~6 this property allows us to describe connections between linear
and quadratic dynamical invariants of the time-dependent Hamiltonian (\re
{GenHam}).
\section{Linear Integrals of Motion}
All invariants of the for
\begin{equation}
P=A\left( t\right) p+B\left( t\right) x+C\left( t\right) ,\quad \frac{dP}{dt
=0, \label{LinInts}
\end{equation
(we call them the Dodonov--Malkin--Man'ko--Trifonov invariants; see, for
example, \cite{Dod:Mal:Man75}, \cite{Dodonov:Man'koFIAN87}, \cit
{Malkin:Man'ko79}, and \cite{Malk:Man:Trif73} and references therein) for
the general variable quadratic Hamiltonian (\ref{GenHam}) can be found in
the following fashion. Use of the differentiation formula (\ref{Diff})
results in the system \cite{Cor-Sot:Sua:SusInv}
\begin{eqnarray}
A^{\prime } &=&2c\left( t\right) A-2a\left( t\right) B, \label{LinCA} \\
B^{\prime } &=&2b\left( t\right) A-2d\left( t\right) B, \label{LinCB} \\
C^{\prime } &=&\left( c\left( t\right) -d\left( t\right) \right) C.
\label{LinCC}
\end{eqnarray
The last equation is explicitly integrated and elimination of $B$ and $A$
from (\ref{LinCA}) and (\ref{LinCB}), respectively, gives the second-order
equations
\begin{eqnarray}
A^{\prime \prime }-\left( \frac{a^{\prime }}{a}+2c-2d\right) A^{\prime
}+4\left( ab-cd+\frac{c}{2}\left( \frac{a^{\prime }}{a}-\frac{c^{\prime }}{c
\right) \right) A &=&0, \label{EqC} \\
B^{\prime \prime }-\left( \frac{b^{\prime }}{b}+2c-2d\right) B^{\prime
}+4\left( ab-cd-\frac{d}{2}\left( \frac{b^{\prime }}{b}-\frac{d^{\prime }}{d
\right) \right) B &=&0. \label{EqBC}
\end{eqnarray
The first is simply the characteristic equation (\ref{in6})--(\ref{in7}).
(It also coincides with the Ehrenfest theorem (\cite{Cor-Sot:Sua:SusInv},
\cite{Ehrenfest}) when $c\leftrightarrow d.)$\medskip
Thus all linear quantum invariants are given b
\begin{equation}
P=A\left( t\right) p+\frac{2c\left( t\right) A\left( t\right) -A^{\prime
}\left( t\right) }{2a\left( t\right) }x+C_{0}\exp \left( \int_{0}^{t}\left(
c\left( s\right) -d\left( s\right) \right) \ ds\right) , \label{GenLinInv}
\end{equation
where $A\left( t\right) $ is a general solution of equation (\ref{EqC})
depending upon two parameters and $C_{0}$ is the third constant. Study of
spectra of the linear dynamical invariants allows one to solve the Cauchy
initial value problem (\cite{Dod:Mal:Man75}, \cite{Dodonov:Man'koFIAN87},
\cite{Malkin:Man'ko79} and \cite{Malk:Man:Trif73}).
\begin{theorem}
\textbf{(Eigenvalue Problem for the Linear Invariants)} I
\begin{equation}
P\left( t\right) =\mu \left( t\right) p+\frac{2c\left( t\right) \mu \left(
t\right) -\mu ^{\prime }\left( t\right) }{2a\left( t\right) }x,
\label{LinMomentum}
\end{equation
then for any solution $A=\mu \left( t\right) $ of the characteristic
equation (\ref{in6})--(\ref{in7}) we hav
\begin{equation}
P\left( t\right) K\left( x,y,t\right) =\beta \left( 0\right) \mu \left(
0\right) \lambda \left( t\right) y\ K\left( x,y,t\right) .
\label{EigenValueProblem}
\end{equation
The eigenfunctions are bounded solutions of the time-dependent Schr\"{o
dinger equation (\ref{Schroedinger}) given b
\begin{equation}
K\left( x,y,t\right) =\frac{1}{\sqrt{2\pi \mu \left( t\right) }}\ e^{i\left(
\alpha \left( t\right) x^{2}+\beta \left( t\right) xy+\gamma \left( t\right)
y^{2}\right) }, \label{KKernel}
\end{equation
where $\lambda \left( t\right) =\exp \left( \int_{0}^{t}\left( c\left(
s\right) -d\left( s\right) \right) \ ds\right) $ an
\begin{eqnarray}
\mu \left( t\right) &=&2\mu \left( 0\right) \mu _{0}\left( t\right) \left(
\alpha \left( 0\right) +\gamma _{0}\left( t\right) \right) , \label{MKernel}
\\
\alpha \left( t\right) &=&\frac{1}{4a\left( t\right) }\frac{\mu ^{\prime
}\left( t\right) }{\mu \left( t\right) }-\frac{c\left( t\right) }{2a\left(
t\right) } \label{AKernel} \\
&=&\alpha _{0}\left( t\right) -\frac{\beta _{0}^{2}\left( t\right) }{4\left(
\alpha \left( 0\right) +\gamma _{0}\left( t\right) \right) }, \notag \\
\beta \left( t\right) &=&-\frac{\beta \left( 0\right) \beta _{0}\left(
t\right) }{2\left( \alpha \left( 0\right) +\gamma _{0}\left( t\right)
\right) } \label{BKernel} \\
&=&\frac{\beta \left( 0\right) \mu \left( 0\right) }{\mu \left( t\right)
\lambda \left( t\right) , \notag \\
\gamma \left( t\right) &=&\gamma \left( 0\right) -\frac{\beta ^{2}\left(
0\right) }{4\left( \alpha \left( 0\right) +\gamma _{0}\left( t\right)
\right) }. \label{CKernel}
\end{eqnarray
(When $\mu =\mu _{0}$ with $\mu _{0}\left( 0\right) =0$ and $\mu
_{0}^{\prime }\left( 0\right) =2a\left( 0\right) \neq 0,$ we obtain the
Green function (\ref{in2})--(\ref{in5}) up to a simple factor.)
\end{theorem}
\begin{proof}
The required solution (\ref{KKernel})--(\ref{CKernel}) has been already
constructed in \cite{Suaz:Sus} and one has to verify formula (\re
{EigenValueProblem}) only. We hav
\begin{eqnarray}
PK\left( x,y,t\right) &=&\left( Ap+Bx\right) K\left( x,y,t\right) \\
&=&\left( \left( 2\alpha A+B\right) x+\beta Ay\right) K\left( x,y,t\right) ,
\notag
\end{eqnarray
wher
\begin{eqnarray}
2\alpha A+B &=&\left( \frac{1}{2a}\frac{\mu ^{\prime }}{\mu }-\frac{c}{a
\right) A+\frac{2cA-A^{\prime }}{2a} \\
&=&\frac{A}{2a}\left( \frac{\mu ^{\prime }}{\mu }-\frac{A^{\prime }}{A
\right) =0 \notag
\end{eqnarray
provided $\mu =A.$ Then $\beta A=\beta \left( 0\right) \mu \left( 0\right)
\lambda $ and the proof is complete.
\end{proof}
Our theorem can be thought as a natural extension to the case of
non-self-adjoint variable quadratic Hamiltonians (\ref{GenHam}) of a
familiar relation between the Green function and linear dynamical invariants
established in \cite{Dod:Mal:Man75}, \cite{Dodonov:Man'koFIAN87}, \cit
{Malkin:Man'ko79} and \cite{Malk:Man:Trif73}. The time-dependent factor in
the eigenvalue (\ref{EigenValueProblem}) corresponds to the statement of
Lemma~2.\medskip\
In this Letter we are interested, among other things, in a direct
verification of Lemma~1 for linear and quadratic dynamical invariants. For
the general Dodonov--Malkin--Man'ko--Trifonov invariant (\ref{GenLinInv}),
without any loss of generality, one can separately consider two cases, say,
when $A\left( t\right) \equiv 0$ with $C_{0}=1$ and when $C_{0}=0.$ I
\begin{equation}
\psi _{1}=\psi \ e^{\int_{0}^{t}\left( c-d\right) ds},
\end{equation
the
\begin{eqnarray}
&&i\frac{\partial \psi _{1}}{\partial t}=i\frac{\partial \psi }{\partial t}\
e^{\int_{0}^{t}\left( c-d\right) ds}+i\left( c-d\right) \psi _{1} \\
&&\qquad \quad =\left( ap^{2}+bx^{2}+cpx+dxp\right) \psi _{1} \notag \\
&&\qquad \qquad +\left( c-d\right) \left( xp-px\right) \psi _{1}=H^{\dagger
}\psi _{1},
\end{eqnarray
which takes care of the first case (we have verified once again the
statement of Lemma~1 for the simplest invariant (\ref{SimplestInvariant
)).\medskip
In the second case, $C_{0}=0,$ we follow Theorem~1 and take a solution
A=\mu \left( t\right) $ of (\ref{EqC}) and (\ref{in6}), which does not have
to satisfy the initial conditions of the Green function (\ref{in8}). Then by
the superposition principle solution of the corresponding Cauchy initial
values problem is given by the integral operator \cite{Suaz:Sus
\begin{equation}
\psi \left( x,t\right) =\int_{-\infty }^{\infty }K\left( x,y,t\right) \ \chi
\left( y\right) \ dy,\qquad \psi \left( x,0\right) =\int_{-\infty }^{\infty
}K\left( x,y,0\right) \ \chi \left( y\right) \ dy \label{KSolution}
\end{equation
with the kernel (\ref{KKernel})--(\ref{CKernel}). Thu
\begin{equation}
\psi _{1}\left( x,t\right) =P\psi \left( x,t\right) =\int_{-\infty }^{\infty
}\left( PK\left( x,y,t\right) \right) \ \chi \left( y\right) \ dy
\label{PKSolution}
\end{equation
(we freely interchange differentiation and integration in this Letter; it
can be justified for certain classes of solutions (\cite{Lieb:Loss}, \cit
{Oh89}, \cite{Per-G:Tor:Mont}, \cite{Velo96})). By choosing $\beta \left(
0\right) \mu \left( 0\right) =1$ in (\ref{EigenValueProblem}) we obtain
\begin{equation}
\psi _{1}\left( x,t\right) =P\psi \left( x,t\right) =e^{\int_{0}^{t}\left(
c-d\right) ds}\ \int_{-\infty }^{\infty }K\left( x,y,t\right) \ \left( y\chi
\left( y\right) \right) \ dy, \label{SecondKSolution}
\end{equation
where the second factor obviously satisfies the Schr\"{o}dinger equation
\ref{Schroedinger}) (see \cite{Suaz:Sus} for details). Repeating the first
step one completes the proof. Eq.~(\ref{SecondKSolution}) provides a
spectral decomposion (\cite{Reed:SimonI}, \cite{Reed:SimonII}) of the linear
invariant $P,$ which has a continuous spectrum.\medskip
Our consideration shows how Lemma~1 works for the general
Dodonov--Malkin--Man'ko--Trifonov invariant --- application of this
invariant to a given solution of the corresponding Cauchy initial value
problem simply produces another solution with the following initial data
\begin{eqnarray}
\psi _{1}\left( x,0\right) &=&P\left( 0\right) \psi \left( x,0\right)
\label{InitialData} \\
&=&\int_{-\infty }^{\infty }K\left( x,y,0\right) \ \left( y\chi \left(
y\right) \right) \ dy. \notag
\end{eqnarray
The reader can easily connect initial conditions of two solutions (\re
{KSolution}) and (\ref{SecondKSolution}) with the help of an analog of the
Fourier transform (see also Ref.~\cite{Sua:Sus} for a formal inversion
operator in general).
\section{Quadratic Dynamical Invariants}
All quantum quadratic integrals of motion are given b
\begin{equation}
E=A\left( t\right) p^{2}+B\left( t\right) x^{2}+C\left( t\right) \left(
px+xp\right) ,\quad \frac{dE}{dt}=0, \label{QuadraticInvariant}
\end{equation
and can be found as follows \cite{Cor-Sot:Sua:SusInv}.
\begin{lemma}
The quadratic dynamical invariants of the Hamiltonian (\ref{GenHam}) have
the for
\begin{eqnarray}
E\left( t\right) &=&\left[ \left( \kappa \ p+\frac{\left( c+d\right) \kappa
-\kappa ^{\prime }}{2a}\ x\right) ^{2}+\frac{C_{0}}{\kappa ^{2}}\ x^{2
\right] \label{InvSymmForm} \\
&&\times \exp \left( \int_{0}^{t}\left( c-d\right) \ ds\right) , \notag
\end{eqnarray
where $C_{0}$ is a constant and function $\kappa \left( t\right) $ is a
solution of the following nonlinear auxiliary equation
\begin{equation}
\kappa ^{\prime \prime }-\frac{a^{\prime }}{a}\kappa ^{\prime }+\left[
4ab+\left( \frac{a^{\prime }}{a}-c-d\right) \left( c+d\right) -c^{\prime
}-d^{\prime }\right] \kappa =C_{0}\frac{\left( 2a\right) ^{2}}{\kappa ^{3}}.
\label{AuxEquation}
\end{equation}
\end{lemma}
The structure of quadratic invariants is once again in perfect agreement
with Lemma~2.
\begin{proof}
In a general form this result has been established in \cit
{Cor-Sot:Sua:SusInv} (see also Refs.~\cite{Lewis:Riesen69}, \cite{Symon70}
and \cite{Yeon:Lee:Um:George:Pandey93} for important earlier articles). A
somewhat different and more direct proof is presented here. It is sufficient
to show that the corresponding linear syste
\begin{eqnarray}
A^{\prime }+4aC-\left( 3c+d\right) A &=&0, \label{EqA} \\
B^{\prime }-4bC+\left( c+3d\right) B &=&0, \label{EqB} \\
C^{\prime }+2\left( aB-bA\right) -\left( c-d\right) C &=&0 \label{EqCOne}
\end{eqnarray
has the following general solutio
\begin{eqnarray}
A\left( t\right) &=&\kappa ^{2}\exp \left( \int_{0}^{t}\left( c-d\right) \
ds\right) , \label{SolA} \\
B\left( t\right) &=&\left[ \left( \frac{\kappa ^{\prime }-\left( c+d\right)
\kappa }{2a}\right) ^{2}+\frac{C_{0}}{\kappa ^{2}}\right] \exp \left(
\int_{0}^{t}\left( c-d\right) \ ds\right) , \label{SolB} \\
C\left( t\right) &=&\kappa \frac{\left( c+d\right) \kappa -\kappa ^{\prime
}{2a}\exp \left( \int_{0}^{t}\left( c-d\right) \ ds\right) , \label{SolC}
\end{eqnarray
where $C_{0}$ is a constant and the function $\kappa \left( t\right) $
satisfies the nonlinear auxiliary equation (\ref{AuxEquation}).\medskip
Indeed the substitutio
\begin{equation}
A\left( t\right) =\widetilde{A}\left( t\right) \lambda \left( t\right)
,\quad B\left( t\right) =\widetilde{B}\left( t\right) \lambda \left(
t\right) ,\quad \quad C\left( t\right) =\widetilde{C}\left( t\right) \lambda
\left( t\right) ,
\end{equation
where $\lambda \left( t\right) =\exp \left( \int_{0}^{t}\left( c-d\right) \
ds\right) ,$ transforms the original system into a more convenient for
\begin{eqnarray}
\widetilde{A}^{\prime }+4a\widetilde{C}-2\left( c+d\right) \widetilde{A}
&=&0, \label{EqAT} \\
\widetilde{B}^{\prime }-4b\widetilde{C}+2\left( c+d\right) \widetilde{B}
&=&0, \label{EqBT} \\
\widetilde{C}^{\prime }+2\left( a\widetilde{B}-b\widetilde{A}\right) &=&0.
\label{EqCT}
\end{eqnarray
Lettin
\begin{equation}
\widetilde{A}=\kappa ^{2}\quad \text{and\quad }\widetilde{A}^{\prime
}=2\kappa \kappa ^{\prime } \label{ATildeKappa}
\end{equation
in the first equation (\ref{EqAT}), we obtai
\begin{eqnarray}
\widetilde{C} &=&\frac{\kappa }{2a}\left( \left( c+d\right) \kappa -\kappa
^{\prime }\right) \label{CTildeKappa} \\
&=&-e^{\int \left( c+d\right) dt}\ \frac{\kappa }{2a}\frac{d}{dt}\left(
\kappa e^{-\int \left( c+d\right) dt}\right) . \notag
\end{eqnarray
Then from the third equation (\ref{EqCT})
\begin{eqnarray}
\widetilde{B} &=&\frac{b}{a}\kappa ^{2}+\frac{1}{2a}\left[ \kappa \frac
\kappa ^{\prime }-\left( c+d\right) \kappa }{2a}\right] ^{\prime }
\label{BTildeKappa} \\
&=&\frac{b}{a}\kappa ^{2}+\frac{1}{2a}\frac{d}{dt}\left[ e^{\int \left(
c+d\right) dt}\ \frac{\kappa }{2a}\frac{d}{dt}\left( \kappa e^{-\int \left(
c+d\right) dt}\right) \right] \notag
\end{eqnarray
and substitution of (\ref{CTildeKappa})--(\ref{BTildeKappa}) into (\ref{EqBT
) give
\begin{equation}
k\frac{d}{dt}\left[ 4abk^{2}\mu ^{2}+k\frac{d}{dt}\left( \mu \left( k\frac
d\mu }{dt}\right) \right) \right] +8abk^{2}\mu \left( k\frac{d\mu }{dt
\right) =0 \label{EqKappa}
\end{equation
in the following temporary notation
\begin{equation}
\mu =\kappa e^{-\int \left( c+d\right) dt},\qquad k=\frac{1}{2a}e^{2\int
\left( c+d\right) dt}. \label{NewNotations}
\end{equation
Usin
\begin{equation}
\mu \left( t\right) =y\left( \tau \right) ,\quad k\frac{d\mu }{dt}=\frac{dy}
d\tau },\quad q=4abk^{2}=\frac{b}{a}e^{4\int \left( c+d\right) dt},
\label{NewNNotations}
\end{equation
one get
\begin{equation}
\frac{d}{d\tau }\left[ qy^{2}+\frac{d}{d\tau }\left( y\frac{dy}{d\tau
\right) \right] +2qy\frac{dy}{d\tau }=0
\end{equation
o
\begin{equation}
y\left( y^{\prime \prime }+qy\right) ^{\prime }+3y^{\prime }\left( y^{\prime
\prime }+qy\right) =0
\end{equation
(see also Ref.~\cite{Lewis:Riesen69} for an important special case). By an
integrating factor
\begin{equation}
\frac{d}{d\tau }\left[ y^{3}\left( y^{\prime \prime }+qy\right) \right]
=0,\quad y^{\prime \prime }+qy=\frac{C_{0}}{y^{3}}, \label{TauIntegral}
\end{equation
and a back substitution with the help o
\begin{equation}
\frac{d^{2}y}{d\tau ^{2}}=\frac{e^{3\int \left( c+d\right) dt}}{\left(
2a\right) ^{2}}\ \left[ \kappa ^{\prime \prime }-\frac{a^{\prime }}{a}\kappa
^{\prime }+\left( \left( \frac{a^{\prime }}{a}-c-d\right) \left( c+d\right)
-c^{\prime }-d^{\prime }\right) \kappa \right]
\end{equation
results in the required first integral of the system
\begin{equation}
\frac{d}{dt}\left[ \frac{\kappa ^{3}}{\left( 2a\right) ^{2}}\left( \kappa
^{\prime \prime }-\frac{a^{\prime }}{a}\kappa ^{\prime }+\left( 4ab+\left(
\frac{a^{\prime }}{a}-c-d\right) \left( c+d\right) -c^{\prime }-d^{\prime
}\right) \kappa \right) \right] =0,
\end{equation
which gives our auxiliary equation (\ref{AuxEquation}).\medskip
The last term in (\ref{BTildeKappa}) can be transformed as follow
\begin{eqnarray*}
&&\frac{1}{2a}\frac{d}{dt}\left[ e^{\int \left( c+d\right) dt}\ \frac{\kappa
}{2a}\frac{d}{dt}\left( \kappa e^{-\int \left( c+d\right) dt}\right) \right]
\\
&&\qquad =e^{-2\int \left( c+d\right) dt}\ \frac{d}{d\tau }\left( y\frac{dy}
d\tau }\right) =e^{-2\int \left( c+d\right) dt}\ \left( yy^{\prime \prime
}+\left( y^{\prime }\right) ^{2}\right) \\
&&\qquad =e^{-2\int \left( c+d\right) dt}\ \left[ \left( \frac{dy}{d\tau
\right) ^{2}-qy^{2}+\frac{C_{0}}{y^{2}}\right] \\
&&\qquad =\left( \frac{\kappa ^{\prime }-\left( c+d\right) \kappa }{2a
\right) ^{2}-\frac{b}{a}\kappa ^{2}+\frac{C_{0}}{\kappa ^{2}},
\end{eqnarray*
in view of (\ref{NewNotations})--(\ref{NewNNotations}) and (\ref{TauIntegral
). We have also utilized a convenient identit
\begin{equation}
\frac{dy}{d\tau }=\frac{1}{2a}\left( \frac{d\kappa }{dt}-\left( c+d\right)
\kappa \right) e^{\int \left( c+d\right) dt}.
\end{equation
Thu
\begin{equation}
\widetilde{B}=\left( \frac{\kappa ^{\prime }-\left( c+d\right) \kappa }{2a
\right) ^{2}+\frac{C_{0}}{\kappa ^{2}} \label{BBTildeKappa}
\end{equation
and the proof is complete.
\end{proof}
The case $a=1/2,$ $b=\omega ^{2}\left( t\right) /2$ and $c=d=0$ corresponds
to the familiar Ermakov--Lewis--Riesenfeld invariant \cite{Ermakov}, \cit
{Lewis67}, \cite{Lewis68}, \cite{Lewis68a}, \cite{Lewis:Riesen69}; see also
Refs.~\cite{Lewis:Leach82a} and \cite{Lewis:Leach82b}. (The corresponding
classical invariant in general is discussed in Refs.~\cite{Symon70} and \cit
{Yeon:Lee:Um:George:Pandey93}.) Many examples of completely integrable
generalized harmonic oscillators and their integrals of motion are given in
Ref.~\cite{Cor-Sot:Sua:SusInv}.\medskip
For a positive constant, $C_{0}>0,$ the quantum dynamical invariant (\re
{InvSymmForm}) can be presented in the standard harmonic oscillator form
\cite{Cor-Sot:Sua:Sus}, \cite{Cor-Sot:Sua:SusInv}, \cite{Reed:SimonI}, \cit
{Reed:SimonII}:
\begin{equation}
E=\frac{\omega \left( t\right) }{2}\left( \widehat{a}\left( t\right)
\widehat{a}^{\dagger }\left( t\right) +\widehat{a}^{\dagger }\left( t\right)
\widehat{a}\left( t\right) \right) , \label{EnOperFactor}
\end{equation
wher
\begin{equation}
\omega \left( t\right) =\omega _{0}\exp \left( \int_{0}^{t}\left( c-d\right)
\ ds\right) ,\qquad \omega _{0}=2\sqrt{C_{0}}>0, \label{omega(t)}
\end{equation
\begin{eqnarray}
\widehat{a}\left( t\right) &=&\left( \frac{\sqrt{\omega _{0}}}{2\kappa }-
\frac{\kappa ^{\prime }-\left( c+d\right) \kappa }{2a\sqrt{\omega _{0}}
\right) x+\frac{\kappa }{\sqrt{\omega _{0}}}\frac{\partial }{\partial x},
\label{a(t)} \\
\widehat{a}^{\dagger }\left( t\right) &=&\left( \frac{\sqrt{\omega _{0}}}
2\kappa }+i\frac{\kappa ^{\prime }-\left( c+d\right) \kappa }{2a\sqrt{\omega
_{0}}}\right) x-\frac{\kappa }{\sqrt{\omega _{0}}}\frac{\partial }{\partial
}, \label{across(t)}
\end{eqnarray
and $\kappa $ is a real-valued solution of the nonlinear auxiliary equation
\ref{AuxEquation}). Here the time-dependent annihilation $\widehat{a}\left(
t\right) $ and creation $\widehat{a}^{\dagger }\left( t\right) $ operators
satisfy the canonical commutation relation
\begin{equation}
\widehat{a}\left( t\right) \widehat{a}^{\dagger }\left( t\right) -\widehat{a
^{\dagger }\left( t\right) \widehat{a}\left( t\right) =1.
\label{commutatora(t)across(t)}
\end{equation
The oscillator-type spectrum of the dynamical invariant $E$ can be obtained
now in a standard way by using the Heisenberg--Weyl algebra of the rasing
and lowering operators (a \textquotedblleft second
quantization\textquotedblright\ \cite{Lewis:Riesen69}, the Fock states)
\begin{equation}
\widehat{a}\left( t\right) \Psi _{n}\left( x,t\right) =\sqrt{n}\ \Psi
_{n-1}\left( x,t\right) ,\quad \widehat{a}^{\dagger }\left( t\right) \Psi
_{n}\left( x,t\right) =\sqrt{n+1}\ \Psi _{n+1}\left( x,t\right) ,
\label{annandcratoperactions}
\end{equation
\begin{equation}
E\left( t\right) \Psi _{n}\left( x,t\right) =\omega \left( t\right) \left( n
\frac{1}{2}\right) \Psi _{n}\left( x,t\right) . \label{Eeigenvp}
\end{equation
The corresponding orthonormal time-dependent eigenfunctions are given b
\begin{equation}
\Psi _{n}\left( x,t\right) =\exp \left( i\frac{\kappa ^{\prime }-\left(
c+d\right) \kappa }{4a\kappa }x^{2}\right) v_{n}\left( \xi \right)
\label{Eeigenfs}
\end{equation
in terms of Hermite polynomials \cite{Ni:Su:Uv}:
\begin{equation}
v_{n}=C_{n}e^{-\xi ^{2}/2}H_{n}\left( \xi \right) ,\qquad \xi =\varepsilon x,
\label{Eeigenfvs}
\end{equation
wher
\begin{equation}
\left\vert C_{n}\right\vert ^{2}=\frac{\varepsilon }{\sqrt{\pi }2^{n}n!
,\qquad \varepsilon ^{2}=\frac{\omega _{0}}{2\kappa ^{2}}=\frac{\sqrt{C_{0}
}{\kappa ^{2}}. \label{Eeigenfsconstant}
\end{equation
Their relation with the original Cauchy initial value problem is discussed
in Section~7 (see Theorem~2). In addition the $n$-dimensional oscillator
wave functions form a basis of the irreducible unitary representation of the
Lie algebra of the noncompact group $SU\left( 1,1\right) $ corresponding to
the discrete positive series $\mathcal{D}_{+}^{j}$ (see \cite{Me:Co:Su},
\cite{Ni:Su:Uv} and \cite{Smir:Shit}).\smallskip\ Operators (\ref{a(t)})--
\ref{across(t)}) allow us to extend these group-theoretical properties to
the general quadratic dynamical invariant (\ref{EnOperFactor}).\medskip
When $C_{0}=0,$ the dynamical invariant (\ref{InvSymmForm}) is given as the
square of a linear invariant up to a simple factor, which resembles the case
of a free particle. If $C_{0}<0,$ one deals with the Hamiltonian of a linear
repulsive \textquotedblleft oscillator\textquotedblright .
\section{Relation Between Linear and Quadratic Invariants}
By Lemma~3 the operators $p^{2},$ $x^{2}$ and $px+xp$ form a basis for all
quadratic invariants. Here we take two linearly independent solutions, say
\mu _{1}=A_{1}$ and $\mu _{2}=A_{2},$ of equations (\ref{in6}) and (\ref{EqC
) and consider the corresponding Dodonov--Malkin--Man'ko--Trifonov
invariants (\ref{GenLinInv}):
\begin{equation}
P_{1}=A_{1}p+B_{1}x,\qquad P_{2}=A_{2}p+B_{2}x.
\end{equation
Introducing the following quadratic invariant
\begin{equation}
E_{1}=P_{1}^{2}\ e^{-\int_{0}^{t}\left( c-d\right) ds},\quad
E_{2}=P_{2}^{2}\ e^{-\int_{0}^{t}\left( c-d\right) ds},\quad E_{3}=\left(
P_{1}P_{2}+P_{2}P_{1}\right) \ e^{-\int_{0}^{t}\left( c-d\right) ds}
\end{equation
with the help of (\ref{dynamicinvariant}) as another basis, one get
\begin{eqnarray}
E &=&C_{1}E_{1}+C_{2}E_{2}+C_{3}E_{3} \label{Linear&Quadratic} \\
&=&\left[ C_{1}P_{1}^{2}+C_{2}P_{2}^{2}+C_{3}\left(
P_{1}P_{2}+P_{2}P_{1}\right) \right] \exp \left( -\int_{0}^{t}\left(
c-d\right) \ ds\right) \notag
\end{eqnarray
for some constants $C_{1},$ $C_{2}$ and $C_{3}.$ As a result the following
operator identity hold
\begin{eqnarray}
&&\left[ \left( \kappa \ p+\frac{\left( c+d\right) \kappa -\kappa ^{\prime
}{2a}\ x\right) ^{2}+\frac{C_{0}}{\kappa ^{2}}\ x^{2}\right] \exp \left(
\int_{0}^{t}\left( c-d\right) \ ds\right) \label{OpIdentity} \\
&&\quad =\left[ C_{1}\left( A_{1}p+B_{1}x\right) ^{2}+C_{2}\left(
A_{2}p+B_{2}x\right) ^{2}\right. \notag \\
&&\qquad \left. +C_{3}\left( \left( A_{1}p+B_{1}x\right) \left(
A_{2}p+B_{2}x\right) +\left( A_{2}p+B_{2}x\right) \left(
A_{1}p+B_{1}x\right) \right) \right] \notag \\
&&\qquad \qquad \times \exp \left( -\int_{0}^{t}\left( c-d\right) \
ds\right) , \notag
\end{eqnarray
wher
\begin{equation}
A_{1}=\mu _{1},\quad B_{1}=\frac{2c\mu _{1}-\mu _{1}^{\prime }}{2a},\qquad
A_{2}=\mu _{2},\quad B_{2}=\frac{2c\mu _{2}-\mu _{2}^{\prime }}{2a}.
\label{OpIdentityAB}
\end{equation
Thus we obtai
\begin{equation}
\kappa ^{2}=\left( C_{1}\mu _{1}^{2}+C_{2}\mu _{2}^{2}+2C_{3}\mu _{1}\mu
_{2}\right) \exp \left( -2\int_{0}^{t}\left( c-d\right) \ ds\right)
\label{NonlinSuperposition}
\end{equation
as a relation between solutions of the nonlinear auxiliary equation (\re
{AuxEquation}) and the linear characteristic equation (\ref{EqC}). In
addition the substitutio
\begin{equation}
\mu _{1}=\kappa _{1}\exp \left( \int_{0}^{t}\left( c-d\right) \ ds\right)
,\qquad \mu _{2}=\kappa _{2}\exp \left( \int_{0}^{t}\left( c-d\right) \
ds\right) \label{MuKappaSubstitution}
\end{equation
transforms the characteristic equation (\ref{EqC}) into our auxiliary
equation (\ref{AuxEquation}) with $C_{0}=0.$ Finally a general solution of
the nonlinear equation is given by the following \textquotedblleft operator
law of cosines\textquotedblright
\begin{equation}
\kappa ^{2}\left( t\right) =C_{1}\kappa _{1}^{2}\left( t\right) +C_{2}\kappa
_{2}^{2}\left( t\right) +2C_{3}\kappa _{1}\left( t\right) \kappa _{2}\left(
t\right) \label{AuxSol}
\end{equation
in terms of two linearly independent solutions $\kappa _{1}$ and $\kappa
_{2} $ of the homogeneous equation. The constant $C_{0}$ is related to the
Wronskian of two linearly independent solutions $\kappa _{1}$ and $\kappa
_{2}:$
\begin{equation}
C_{1}C_{2}-C_{3}^{2}=C_{0}\frac{\left( 2a\right) ^{2}}{W^{2}\left( \kappa
_{1},\kappa _{2}\right) },\quad W\left( \kappa _{1},\kappa _{2}\right)
=\kappa _{1}\kappa _{2}^{\prime }-\kappa _{1}^{\prime }\kappa _{2}
\label{AuxSolWronskian}
\end{equation
(more details are given in Appendix~A). This is a well-known nonlinear
superposition property of the so-called Ermakov systems (see, for example,
\cite{Corant:Snyder58}, \cite{Eliezer:Gray76}, \cite{Ermakov}, \cit
{Leach:Andrio08}, \cite{Leach:Andriopo08}, \cite{Leach:Kar:Nuc:Andrio05},
\cite{Lewis68a}, \cite{Mah:Leach07}, \cite{Nucci:Leach05}, \cit
{PadillaMaster}, \cite{Pinney50}, \cite{Schuch08} and references therein).
Here we have obtained this \textquotedblleft nonlinear superposition
principle\textquotedblright\ (or Pinney's solution) in an operator form by
multiplication and addition of the linear dynamical invariants together with
an independent characterization of all quantum quadratic invariants, which
seems to be missing, in general, in the available literature (see also \cit
{Lewis67} and \cite{Lewis68a} for an important classical case). An extension
is given in the last section.\medskip
It is worth noting, in conclusion, that the linear invariants of Dodonov,
Malkin, Man'ko and Trifonov (\cite{Dod:Mal:Man75}, \cit
{Dodonov:Man'koFIAN87}, \cite{Malkin:Man'ko79} and \cite{Malk:Man:Trif73})
can be presented as follow
\begin{eqnarray}
P_{1} &=&\left( \kappa _{1}p+\frac{\left( c+d\right) \kappa _{1}-\kappa
_{1}^{\prime }}{2a}x\right) \exp \left( \int_{0}^{t}\left( c-d\right) \
ds\right) , \\
P_{2} &=&\left( \kappa _{2}p+\frac{\left( c+d\right) \kappa _{2}-\kappa
_{2}^{\prime }}{2a}x\right) \exp \left( \int_{0}^{t}\left( c-d\right) \
ds\right)
\end{eqnarray
in terms of two linearly independent solutions, $\kappa _{1}$ and $\kappa
_{2},$ of the homogeneous equation (\ref{AuxEquation}) when $C_{0}=0.$
Comparing these expressions with the form of the quadratic invariant (\re
{InvSymmForm}) at $C_{0}=0$ (no \textquotedblleft
potential\textquotedblright , a \textquotedblleft free
particle\textquotedblright ), when the operator square is complete, one can
treat the linear invariants as \textquotedblleft operator square
roots\textquotedblright\ \cite{Deb:Mikus} of the special quadratic
invariants (see also Lemma~2 regarding a convenient common factor).
\section{Quadratic Invariants and Cauchy Initial Value Problem}
Our decomposition (\ref{Linear&Quadratic}) of the quantum quadratic
invariant in terms of products of the linear ones not only results in the
Pinney solution (\ref{AuxSol})--(\ref{AuxSolWronskian}) of the corresponding
generalized Ermakov system (\ref{AuxEquation}) in a form of an
\textquotedblleft operator law of cosines\textquotedblright , but also
provides a somewhat better understanding, with the help of Lemma~1 and
properties of the linear invariants discussed in Section~4, how the
quadratic invariants act on solutions of the original time-dependent Sch
\"{o}dinger equation. Indeed by (\ref{KSolution}) for two different
solutions, say $A_{1}=\mu _{1}$ and $A_{2}=\mu _{2},$ of the characteristic
equation, (\ref{EqC}), we have in an operator form
\begin{equation}
\psi \left( x,t\right) =K_{1}\left( t\right) \left[ \chi _{1}\left( y\right)
\right] =K_{2}\left( t\right) \left[ \chi _{2}\left( y\right) \right]
\label{K12Solution}
\end{equation
in view of uniqueness of the Cauchy initial value problem (see, for example,
Refs.~\cite{Cor-Sot:Sua:SusInv} and \cite{Suaz:Sus}). Then by (\re
{SecondKSolution})
\begin{equation}
P_{1}\psi =e^{\int_{0}^{t}\left( c-d\right) ds}\ K_{1}\left( y\chi
_{1}\right) ,\quad P_{2}\psi =e^{\int_{0}^{t}\left( c-d\right) ds}\
K_{2}\left( y\chi _{2}\right) \label{P12Act}
\end{equation
an
\begin{eqnarray}
E\psi &=&e^{-\int_{0}^{t}\left( c-d\right) ds}\ \left[ C_{1}P_{1}^{2}\psi
+C_{2}P_{2}^{2}\psi +C_{3}\left( P_{1}P_{2}+P_{2}P_{1}\right) \psi \right]
\label{EAct} \\
&=&e^{\int_{0}^{t}\left( c-d\right) ds}\ \left[ C_{1}K_{1}\left( y^{2}\chi
_{1}\right) +C_{2}K_{2}\left( y^{2}\chi _{2}\right) \right] \notag \\
&&+C_{3}\left[ P_{1}\left( K_{2}\left( y\chi _{2}\right) \right)
+P_{2}\left( K_{1}\left( y\chi _{1}\right) \right) \right] . \notag
\end{eqnarray
We hav
\begin{eqnarray*}
&&K_{2}\left( y\chi _{2}\right) =K_{1}\left( \varphi _{1}\right) ,\qquad
\varphi _{1}=K_{1}^{-1}\left( 0\right) \left[ K_{2}\left( 0\right) \left(
y\chi _{2}\right) \right] , \\
&&K_{1}\left( y\chi _{1}\right) =K_{2}\left( \varphi _{2}\right) ,\qquad
\varphi _{2}=K_{2}^{-1}\left( 0\right) \left[ K_{1}\left( 0\right) \left(
y\chi _{1}\right) \right]
\end{eqnarray*
with the help of an analog of the Fourier transform. Therefor
\begin{equation*}
P_{1}\left( K_{2}\left( y\chi _{2}\right) \right) =e^{\int_{0}^{t}\left(
c-d\right) ds}\ K_{1}\left( y\varphi _{1}\right) ,\quad P_{2}\left(
K_{1}\left( y\chi _{1}\right) \right) =e^{\int_{0}^{t}\left( c-d\right) ds}\
K_{2}\left( y\varphi _{2}\right) ,
\end{equation*
and finall
\begin{equation}
E\psi =e^{\int_{0}^{t}\left( c-d\right) ds}\ \left[ C_{1}K_{1}\left(
y^{2}\chi _{1}\right) +C_{2}K_{2}\left( y^{2}\chi _{2}\right) +C_{3}\left(
K_{1}\left( y\varphi _{1}\right) +K_{2}\left( y\varphi _{2}\right) \right)
\right] , \label{ESolution}
\end{equation
where each term satisfies the corresponding Schr\"{o}dinger equation. By the
superposition principle we arrive at a new solution in a complete agreement
with our Lemma~1. The corresponding initial data follow from (\ref{ESolution
) at $t=0$ and the time-evolution operator (\ref{CauchyInVProb}) can be
applied (see also Eq.~(\ref{SpectralDecomposition}) for an eigenfunction
expansion).\medskip
On the second hand one can always expand a square-integrable solution (\re
{KSolution}) of the Cauchy initial value problem in the standard for
\begin{eqnarray}
\psi \left( x,t\right) &=&\int_{-\infty }^{\infty }K\left( x,y,t\right) \
\chi \left( y\right) \ dy \label{KEexpansion} \\
&=&\sum_{n=0}^{\infty }c_{n}\left( t\right) \ \Psi _{n}\left( x,t\right) ,
\notag
\end{eqnarray
wher
\begin{equation}
c_{n}\left( t\right) =\int_{-\infty }^{\infty }\Psi _{n}^{\ast }\left(
x,t\right) \psi \left( x,t\right) \ dx \label{KEexpansionConst}
\end{equation
(we use the asterisk for complex conjugate) by the Riesz--Fisher theorem
\cite{Reed:SimonI}, \cite{Reed:SimonII}, \cite{Rudin} due to completeness of
the eigenfunctions (\ref{Eeigenfs})--(\ref{Eeigenfsconstant}) at all times.
Then by the Fubini theorem
\begin{equation}
c_{n}\left( t\right) =\int_{-\infty }^{\infty }\chi \left( y\right) \left(
\int_{-\infty }^{\infty }\Psi _{n}^{\ast }\left( x,t\right) K\left(
x,y,t\right) \ \ dx\right) \ dy \label{KEexpansionConstant}
\end{equation
and the second integral can be evaluated with the help of Eqs.~(\ref{KKernel
), (\ref{Eeigenfs})--(\ref{Eeigenfsconstant}) and (\ref{Erd}) as follow
\begin{eqnarray}
&&\int_{-\infty }^{\infty }\Psi _{n}^{\ast }\left( x,t\right) K\left(
x,y,t\right) \ \ dx \label{KPsiIntegral} \\
&&\ =\frac{i^{n}e^{-i\left( n+1/2\right) \varphi }}{\left( \sqrt{\pi
2^{n}n!\right) ^{1/2}}\left( \frac{\sqrt{C_{0}}}{C_{0}\left( \frac{\kappa
_{1}}{\kappa }\right) ^{2}+\left( \frac{\kappa _{1}\kappa ^{\prime }-\kappa
_{1}^{\prime }\kappa }{2a}\right) ^{2}}\right) ^{1/4}\exp \left( \frac{1}{2
\int_{0}^{t}\left( d-c\right) \ ds\right) \notag \\
&&\quad \times \exp \left( i\left( \gamma +\frac{\beta ^{2}\left( 0\right)
\kappa _{1}^{2}\left( 0\right) }{4a\kappa _{1}}\frac{\kappa \left( \kappa
_{1}\kappa ^{\prime }-\kappa _{1}^{\prime }\kappa \right) }{C_{0}\left(
\frac{\kappa _{1}}{\kappa }\right) ^{2}+\left( \frac{\kappa _{1}\kappa
^{\prime }-\kappa _{1}^{\prime }\kappa }{2a}\right) ^{2}}\right) y^{2}\right)
\notag \\
&&\quad \times \exp \left( -\frac{\beta ^{2}\left( 0\right) \kappa
_{1}^{2}\left( 0\right) \sqrt{C_{0}}}{C_{0}\left( \frac{\kappa _{1}}{\kappa
\right) ^{2}+\left( \frac{\kappa _{1}\kappa ^{\prime }-\kappa _{1}^{\prime
}\kappa }{2a}\right) ^{2}}\frac{y^{2}}{2}\right) H_{n}\left( \frac{\beta
\left( 0\right) \kappa _{1}\left( 0\right) C_{0}^{1/4}y}{\sqrt{C_{0}\left(
\frac{\kappa _{1}}{\kappa }\right) ^{2}+\left( \frac{\kappa _{1}\kappa
^{\prime }-\kappa _{1}^{\prime }\kappa }{2a}\right) ^{2}}}\right) . \notag
\end{eqnarray
Here $\kappa _{1}=\mu \exp \left( \int_{0}^{t}\left( d-c\right) \ ds\right) $
and $\kappa $ are the corresponding solutions of auxiliary equation (\re
{AuxEquation}) with $C_{0}=0$ and $C_{0}\neq 0,$ respectively, with the
Wronskian $W\left( \kappa _{1},\kappa \right) =\kappa _{1}\kappa ^{\prime
}-\kappa _{1}^{\prime }\kappa $ and
\begin{equation}
\tan \varphi =\frac{\kappa }{\kappa _{1}}\frac{\kappa _{1}\kappa ^{\prime
}-\kappa _{1}^{\prime }\kappa }{2a\sqrt{C_{0}}}, \label{tangent}
\end{equation
\begin{equation}
C_{0}\left( \frac{\kappa _{1}}{\kappa }\right) ^{2}+\left( \frac{\kappa
_{1}\kappa ^{\prime }-\kappa _{1}^{\prime }\kappa }{2a}\right) ^{2}=constant
\label{constant}
\end{equation
by (\ref{A4}). (The computational details are left to the reader; our
equation (\ref{constant}) is equivalent to the classical Ermakov invariant
\cite{Ermakov}.) We may choose $\beta \left( 0\right) \kappa _{1}\left(
0\right) =1$ and arrive at the following result.
\begin{theorem}
\textbf{(Eigenfunction Expansions)} Solution of the Cauchy initial value
problem (\ref{KSolution}) in $L^{2}\left(
\mathbb{R}
\right) $ can be obtained as an infinite series of multiples of the
quadratic invariant eigenfunctions (\ref{Eeigenfs})
\begin{equation}
\psi \left( x,t\right) =\sum_{n=0}^{\infty }c_{n}\left( t\right) \ \Psi
_{n}\left( x,t\right) , \label{QIExpSolution}
\end{equation
where the time-dependent coefficients are given b
\begin{eqnarray}
c_{n}\left( t\right) &=&i^{n}\left( \frac{\delta }{\sqrt{\pi }2^{n}n!
\right) ^{1/2}e^{-i\left( n+1/2\right) \varphi }\exp \left( \frac{1}{2
\int_{0}^{t}\left( d-c\right) \ ds\right) \label{c(t)constant} \\
&&\times \int_{-\infty }^{\infty }\exp \left( i\xi y^{2}\right) e^{-\delta
^{2}y^{2}/2}H_{n}\left( \delta y\right) \chi \left( y\right) \ dy. \notag
\end{eqnarray
Here $\kappa _{1}\left( t\right) $and $\kappa \left( t\right) $ are
real-valued solutions of the homogeneous and nonhomogeneous auxiliary
equations (\ref{AuxEquation}), respectively, with the Wronskian $W\left(
t\right) =\kappa _{1}\kappa ^{\prime }-\kappa _{1}^{\prime }\kappa .$ The
phases $\varphi \left( t\right) $ and $\gamma \left( t\right) $ are
determined in terms of these solutions as follow
\begin{equation}
\varphi =\arctan \left( \frac{\kappa }{\kappa _{1}}\frac{\kappa _{1}\kappa
^{\prime }-\kappa _{1}^{\prime }\kappa }{2a\sqrt{C_{0}}}\right) ,\quad \frac
d\varphi }{dt}=2\sqrt{C_{0}}\frac{a}{\kappa ^{2}}, \label{phase}
\end{equation
\begin{equation}
\frac{d\gamma }{dt}+\frac{a}{\kappa _{1}^{2}}=0 \label{gamma(t)}
\end{equation
an
\begin{equation}
\delta =\frac{C_{0}^{1/4}}{\sqrt{C_{0}\left( \frac{\kappa _{1}}{\kappa
\right) ^{2}+\left( \frac{\kappa _{1}\kappa ^{\prime }-\kappa _{1}^{\prime
}\kappa }{2a}\right) ^{2}}}>0, \label{deltaconstant}
\end{equation
\begin{equation}
\xi =\gamma +\frac{\delta ^{2}}{2\sqrt{C_{0}}}\frac{\left( \kappa _{1}\kappa
^{\prime }-\kappa _{1}^{\prime }\kappa \right) \kappa }{2a\kappa _{1}}
\label{xiconstant}
\end{equation
are constants. A spectral decomposition of the quadratic invariant $E$ in
the space of $L^{2}$-solutions is given b
\begin{equation}
E\left( t\right) \psi \left( x,t\right) =\omega \left( t\right)
\sum_{n=0}^{\infty }c_{n}\left( t\right) \left( n+\frac{1}{2}\right) \ \Psi
_{n}\left( x,t\right) , \label{SpectralDecomposition}
\end{equation
where the \textquotedblleft frequency\textquotedblright\ $\omega \left(
t\right) $ is defined by (\ref{omega(t)}).
\end{theorem}
It is worth noting tha
\begin{equation}
\frac{d}{dt}\left[ \gamma +\frac{\delta ^{2}}{2\sqrt{C_{0}}}\frac{\kappa }
\kappa _{1}}\left( \frac{W}{2a}\right) \right] =0 \label{PhaseConstant}
\end{equation
by (\ref{gamma(t)}) and (\ref{A6}), which means that the phase factor (\re
{xiconstant}) in front of $y^{2}$ in the second integral of Eq.~(\re
{c(t)constant}) is indeed a constant. The second equation (\ref{phase})
follows from (\ref{PhaseConstant}) with the help of (\ref{tangent}) and (\re
{gamma(t)}).\medskip
Finally by choosing in Eqs.~(\ref{QIExpSolution})--(\ref{c(t)constant}) a
special (square) integrable initial data of the for
\begin{equation}
\chi _{m}\left( y\right) =\exp \left( -i\xi y^{2}\right) e^{-\delta
^{2}y^{2}/2}H_{m}\left( \delta y\right) \label{chispecial}
\end{equation
we conclude in view of the orthogonality property of Hermite polynomials
that the time-dependent wave function
\begin{eqnarray}
\psi _{m}\left( x,t\right) &=&D_{m}\exp \left( \frac{1}{2}\int_{0}^{t}\left(
d-c\right) \ ds\right) \label{psispecial} \\
&&\times e^{-i\left( m+1/2\right) \varphi \left( t\right) }\ \Psi _{m}\left(
x,t\right) \notag
\end{eqnarray
are particular solutions of the Schr\"{o}dinger equation (\ref{Schroedinger
)--(\ref{GenHam}) for arbitrary constants $D_{m}:
\begin{equation}
i\frac{\partial \psi _{m}}{\partial t}=H\psi _{m},\qquad E\psi _{m}=\omega
\left( m+\frac{1}{2}\right) \psi _{m}. \label{Schroedingereigenfunctions}
\end{equation
They are also eigenfunctions of the quadratic invariant $E$\ corresponding
to the discrete \textquotedblleft spectrum\textquotedblright
\begin{equation}
\left\langle E\right\rangle _{m}=\int_{-\infty }^{\infty }\psi _{m}^{\ast
}E\psi _{m}\ dx=\left\vert D_{m}\right\vert ^{2}\omega _{0}\left( m+\frac{1}
2}\right) , \label{psispectrum}
\end{equation
see (\ref{Eeigenvp}). (It can be verified by direct substitution; the
details are left to the reader.) The explicit wave functions (\re
{psispecial}) are derived here without separation of the variables with the
aid of our Theorem~2 and certain variants of the Ermakov invariant. If
\left\vert D_{m}\right\vert =1,$ solution (\ref{QIExpSolution}) takes the
for
\begin{equation}
\psi \left( x,t\right) =\sum_{n=0}^{\infty }i^{n}e^{-i\left( n+1/2\right)
\varphi \left( 0\right) }\ \psi _{n}\left( x,t\right) \int_{-\infty
}^{\infty }\psi _{n}^{\ast }\left( y,0\right) \chi \left( y\right) \ dy
\label{wavefunctionexpan}
\end{equation
in terms of the time-dependent wave functions (\ref{psispecial}) provided
tha
\begin{equation}
\varepsilon \left( 0\right) =\frac{C_{0}^{1/4}}{\kappa \left( 0\right)
=\delta ,\quad \frac{\left( c\left( 0\right) +d\left( 0\right) \right)
\kappa \left( 0\right) -\kappa ^{\prime }\left( 0\right) }{4a\left( 0\right)
\kappa \left( 0\right) }=\xi \label{inconsts}
\end{equation
(see also \cite{Yeon:Lee:Um:George:Pandey93} for the path-integral method).
The traditional operator approach (for the parametric oscillator) is
presented in \cite{Lewis:Riesen69} and/or elsewhere (see also \cit
{Andrio:Leach05} and \cite{Leach90}).
\section{A General Nonlinear Superposition Principle for Ermakov's Equations}
The Pinney superposition formula (\ref{AuxSol})--(\ref{AuxSolWronskian})
allows one to construct solutions of the nonlinear auxiliary equation (\re
{AuxEquation}) in terms of given solutions of the corresponding linear
equation. In general we take two different solutions, say $\kappa _{1}$ and
\kappa _{2},$ of the generalized Ermakov equations (\ref{AuxEquation}) with
C_{0}^{\left( 1,2\right) }\neq 0$ and consider two quadratic invariants
\begin{eqnarray}
E_{1}\left( t\right) &=&\left[ \left( \kappa _{1}\ p+\frac{\left( c+d\right)
\kappa _{1}-\kappa _{1}^{\prime }}{2a}\ x\right) ^{2}+\frac{C_{0}^{\left(
1\right) }}{\kappa _{1}^{2}}\ x^{2}\right] \lambda \left( t\right) , \\
E_{2}\left( t\right) &=&\left[ \left( \kappa _{2}\ p+\frac{\left( c+d\right)
\kappa _{2}-\kappa _{2}^{\prime }}{2a}\ x\right) ^{2}+\frac{C_{0}^{\left(
2\right) }}{\kappa _{2}^{2}}\ x^{2}\right] \lambda \left( t\right) .
\end{eqnarray
Their arbitrary linear combination
\begin{equation}
D_{1}E_{1}\left( t\right) +D_{2}E_{2}\left( t\right) =E\left( t\right)
\end{equation
($D_{1}$ and $D_{2}$ are constants), is also a quadratic invariant given by
\ref{InvSymmForm}) for a certain solution $\kappa $ of the nonlinear
auxiliary equation (\ref{AuxEquation}). Thus the following operator identity
hold
\begin{eqnarray}
&&\left( \kappa \ p+\frac{\left( c+d\right) \kappa -\kappa ^{\prime }}{2a}\
x\right) ^{2}+\frac{C_{0}}{\kappa ^{2}}\ x^{2}
\label{QuadraticInvariantIdentity} \\
&&\qquad =D_{1}\left[ \left( \kappa _{1}\ p+\frac{\left( c+d\right) \kappa
_{1}-\kappa _{1}^{\prime }}{2a}\ x\right) ^{2}+\frac{C_{0}^{\left( 1\right)
}{\kappa _{1}^{2}}\ x^{2}\right] \notag \\
&&\qquad \quad +D_{2}\left[ \left( \kappa _{2}\ p+\frac{\left( c+d\right)
\kappa _{2}-\kappa _{2}^{\prime }}{2a}\ x\right) ^{2}+\frac{C_{0}^{\left(
2\right) }}{\kappa _{2}^{2}}\ x^{2}\right] \notag
\end{eqnarray
and in a similar fashion we arrive at a general nonlinear superposition
principle for the solutions of Ermakov's equations (\ref{AuxEquation})
\begin{equation}
\kappa ^{2}\left( t\right) =D_{1}\kappa _{1}^{2}\left( t\right) +D_{2}\kappa
_{2}^{2}\left( t\right) , \label{GenNonLinSupPrinciple}
\end{equation
where the constant $C_{0}$ is given b
\begin{equation}
C_{0}-C_{0}^{\left( 1\right) }D_{1}^{2}-C_{0}^{\left( 2\right)
}D_{2}^{2}=D_{1}D_{2}\left[ \frac{W^{2}\left( \kappa _{1},\kappa _{2}\right)
}{\left( 2a\right) ^{2}}+C_{0}^{\left( 1\right) }\frac{\kappa _{2}^{2}}
\kappa _{1}^{2}}+C_{0}^{\left( 2\right) }\frac{\kappa _{1}^{2}}{\kappa
_{2}^{2}}\right] \text{\label{Constant}}
\end{equation
with $W\left( \kappa _{1},\kappa _{2}\right) $ being the Wronskian of two
solutions (see Appendix~A). One can also derive this property by adding the
corresponding solutions (\ref{SolA})--(\ref{SolC}) of the original linear
system (\ref{EqA})--(\ref{EqCOne}) or with the help of the Pinney formula
\ref{AuxSol})--(\ref{AuxSolWronskian}). The details are left to the
reader.\medskip
\noindent \textbf{Acknowledgments.\/} The author thanks George E. Andrews,
Carlos Castillo-Ch\'{a}vez, Victor V. Dodonov, Vladimir~I. Man'ko, Svetlana
Roudenko and Kurt Bernardo Wolf for support, valuable comments and
encouragement. I am grateful to Peter~G.~L.~Leach for careful reading of the
manuscript --- his numerous suggestions have helped to improve the
presentation.
|
\section{Introduction}
Observations of the Galactic Center provide strong evidence for the
existence of a supermassive black hole in Sgr A* (which hereafter refers
to the black hole, the accretion flow, and the radio source). Sgr A*'s
proximity allows us to perform observations with higher angular
resolution than other galactic nuclei. Estimates of Sgr A*'s mass
$M=4.5 \pm 0.4 \times 10^6 M_{\odot}$ and distance $D = 8.4 \pm 0.4$ kpc
(\citealt{ghez:2008}, \citealt{gillessen:2009}) indicate that it has the
largest angular size of any known black hole ($G M/(c^2 D) \simeq 5.3
{\rm \mu as}$).
Recent 230 GHz VLBI constrains the structure of Sgr A* on angular scales
comparable to the size of the black hole horizon
(\citealt{doeleman:2008}, see also Doeleman contribution to this
conference proceeding). Using a two-parameter, symmetric Gaussian
brightness distribution model VLBI infers a full width at half maximum
FWHM $= 37^{+16}_{-10}$ ${\rm \mu as}$. This is smaller than the
apparent diameter of the black hole: $\approx 2 \sqrt{27} G M /(c^2 D)
\simeq 55 {\rm \mu as}$ (this depends only weakly on black hole spin).
Sgr~A* radio - submm emission is usually modeled as synchrotron
emission, with the turnover at $\sim 230$ GHz indicating transition from
optically thick to optically thin emission. The model of accretion and
its geometry is still under debate because Sgr A* is dim at shorter
wavelengths (in NIR and X-ray band) or completely obscured (UV and
optical light). Moreover in the NIR and X-ray the source is not resolved
and we only have upper limits for its quiescent luminosity.
Sgr~A* is a strongly sub-Eddington source ($L_{Bol} \approx 10^{-9}
L_{Edd}$). Models suggest that the source is accreting inefficiently
(in the sense that $L_{bol}/(\dot{M} c^2) \ll 1$), which justifies
treating the dynamics of the flow and its radiative properties
independently. Therefore most of Sgr~A* models published to this day
consists of two separate parts: plasma dynamics model and/or radiative
transfer model. We can further categorize the dynamical models into:
accretion flow vs. outflow models, stationary vs. time-dependent,
Newtonian/post-Newtonian vs. General Relativistic, models covering small
region (a couple of gravitational radii) vs. large region (thousands of
gravitational radii). The radiative transfer modeling usually uses ray
tracing (always relativistic) or Monte Carlo methods (nonrelativistic as
well as fully relativistic). Ray tracing allows one to model source
images at sub-mm frequencies and the SED of direct synchrotron emission
(radio,sub-mm). Monte Carlo allows one to model the Compton scattering
and multiwavelength (from radio to gamma-rays) SED of Sgr~A*. In
Table~\ref{table_models} we summarize the recent progress in models.
\begin{table}[!ht]
\smallskip
\begin{center}
{\small
\begin{tabular}{lcccc}
\tableline
Reference & dynamical & radiative & plasma& range \\
& model & model & & of model \\
\tableline
\citet{narayan:1998} & stat. rel. ADAF & non-rel. MC & th & $10^5 R_g$\\
\citet{markoff:2001}& Jet& scaling & non-th & --\\
\citet{yuan:2003}& stat non-rel. RIAF & non-rel rays & th+non-th& $2 \times 10^5 R_g$\\
\citet{ohsuga:2005}& MHD-time dep. & non-rel. MC & th & $60 R_g$\\
\citet{goldston:2005}& MHD-time dep. & polarized non-rel. rays & th+non-th & $512 R_g$\\
\citet{broderick:2006b}& stat. non-rel RIAF& polarized RT & non-th & $2 \times 10^5 R_g$\\
\citet{moscibrodzka:2007}& MHD-time dep. & non-rel. MC & th+non-th & $2.4 \times 10^3 R_g$\\
\citet{loeb:2007}& Jet & scaling & th+non-th & --\\
\citet{huang:2007}& stat. RIAF& RT & th & $2 \times 10^5 R_g$ \\
\citet{markoff:2007}& Jet& non-rel rays /w corr & non-th & --\\
\citet{huang:2009}& stat. rel. RIAF & RT & th & $10^4 R_g$\\
\citet{broderick:2009}& stat. rel. RIAF & RT & th+non-th & $2 \times 10^5 R_g$ \\
\citet{chan:2009}& MHD-time dep.& non-rel rays /w corr. & th+non-th& $43 R_g$\\
\citet{yuan:2009}& stat. rel. RIAF& RT& th & $100 R_g$\\
\citet{hilburn:2009}& GRMHD-time dep. & non-rel MC & th & $40 R_g$\\
\citet{dexter:2009}& GRMHD-time dep. & RT & th & $40 R_g$\\
\citet{moscibrodzka:2009}& GRMHD-time dep. & RT + rel. MC & th & $40R_g$\\
\tableline
\end{tabular}
}
\caption{Summary of selected models of Sgr~A*. Abbreviations: RT-ray tracing,
MC-Monte Carlo, GR-general relativistic, RIAF-radiatively inefficient
accretion flow, ADAF-advection dominate accretion flow, plasma-particles
distribution,
th-thermal, non-th-non-thermal, range-model radial range.
}\label{table_models}
\end{center}
\end{table}
\section{Methodology}
Our numerical model of accretion onto Sgr A* consist of: a physical
model of the accretion flow dynamics with its numerical realization, and
a radiative transfer model. We assume that the accreting plasma is
geometrically thick, optically thin, turbulent and it accretes onto a
spinning black hole. The spin angular momentum $J$ of the black hole,
whose magnitude is parameterized by $a_* = J c/(G M^2)$, is assumed to
be aligned with the angular momentum of the accretion flow.
The accretion flow in Sgr~A* is collisionless. We assume that ions and
electrons have thermal distributions, possibly with different
temperatures. In our model we allow the electron temperature $T_e$ to
differ from the ion temperature $T_i$, but we fix the ratio
($T_i/T_e=const$). The plasma equation of state is described by the
$P=e(\gamma-1)$ relation with $\gamma=13/9$ (non-relativistic ions and
relativistic electrons). A more physical model would evolve $T_e$ and
$T_i$ independently with a model for dissipation and energy exchange
between the electrons and ions, but this would complicate the model and
introduce a host of new parameters.
We realize the physical model using an axisymmetric version of the GRMHD
code {\tt harm} (\citealt{gammie:2003}). As initial conditions we adopt
an analytical model of a thick disk (a torus) in hydrostatic equilibrium
\citep{fishbone:1976}. Since the code is not designed to evolve MHD
equations in vacuum, we surround the equilibrium torus with a hot, low
density plasma which does not influence the torus evolution. We seed
the torus with a poloidal, concentric loops of weak magnetic field.
Small perturbations are added to the internal energy which allows to for
magnetorotational instability and turbulence development. We solve the
GRMHD evolution equations of evolution until a quasi-equilibrium
accretion flow is established (meaning that the flow is not evolving on
the dynamical timescale). The details of torus initial setup and
discussion of the flow evolution is presented in \citet{mckinney:2004}.
Our numerical domain extends from the black hole event horizon to $40
GM/c^2 = 1.8$ AU or $210 \mu as$, but is only in equilibrium to $\sim 15
GM/c^2 = 0.7$AU or $80 \mu as$. Since low frequency emission from Sgr
A* is believed to arise at larger radius, we are unable to model the low
frequency (radio, mm) portion of spectral energy distribution (SED).
Our (untested) hypothesis is that at $r < 15 GM/c^2$ the model
accurately represent the inner portions of a relaxed accretion flow
extending over many decades in radius.
Our fluid dynamical model is scale free but the radiative transfer
calculation is not. Therefore we need to specify the simulation length
unit ${\mathcal L}=GM/c^2$, time unit ${\mathcal T}=GM/c^3$, and mass
unit ${\mathcal M}$ (which scales the mass accretion rate). ${\mathcal
L}$ and ${\mathcal T}$ are defined by the black hole mass adopted from
observations. ${\mathcal M}$ is a free parameter set to reproduce the
submillimeter flux $F_{230GHz}=3.4 Jy$, \citep{marrone:2006b}. To
``observe'' the numerical model we need to specify the inclination $i$
of the black hole spin to the line of sight. The distance to the
observer is assumed to be $D=8.4 kpc$.
The SED is calculated in a Monte Carlo fashion. We use {\tt grmonty} - a
general relativistic radiative transfer scheme (\citealt{dolence:2009}) to
calculate the light propagation in the strong gravity taking into account
synchrotron emission and absorption and Compton scattering. {\tt grmonty} has
been extensively tested. As one of our tests we compared the code performance
in a flat space with independent Monte Carlo scheme {\tt Sphere} (kindly
provided by its author, Shane Davies). In Fig.\ref{test_fig} we present a
standard test for Compton scattering in the spherical, homogenous, cloud of
hot plasma \citep{pss:1979}.
\begin{figure}[!ht]
\begin{center}
\includegraphics[scale = 0.5, angle = 0]{proc_fig1.eps}
\end{center}
\caption{Code testing: {\tt grmonty} (points) and {\tt Sphere} (independent
Monte Carlo code for radiative transfer, solid line) SED comparison. In this
test problem a thermal point source emits photons from the center of the
spherical, hot plasma cloud. The cloud is characterized by two parameters:
optical thickness $\tau$ (here $\tau=0.1$) and electron temperature
$\Theta_e$ ($\Theta_e=kT_e/m_ec^2=4.$). Lower panel shows the difference
between two SEDs.
}\label{test_fig}
\end{figure}
In the present models we compute spectra through a time slice of a
simulation as if the light had infinite speed ('fast light'
approximation). In reality the light crossing time is comparable to the
dynamical time which suggest radiative transfer through the changing
medium ($t_{dyn} \sim t_{cross} \sim 20$s in Sgr A*). Finally we use a
ray-tracing scheme, {\tt ibothros} (\citealt{noble:2007}), which
accounts only for synchrotron emission and absorption in order to
calculate the 230 GHz intensity maps of an accretion flow as seen by a
distant observer (again in the fast light approximation). To estimate
the size of the emitting region we calculate the eigenvalues of the
matrix formed by taking the second angular moments of the image one the
sky (the lengths of the principal axes). The eigenvalues along the major
and minor axis are then compared to the FWHM VLBI constraint.
We accept or reject numerical models based on several observational
constraints: the flux at 230 GHz (we scale our models to reconstruct the
flux F=3.4 Jy at 230 GHz), submillimeter spectral slope $\alpha$ around
a turn over frequency (230-690 GHz, $\alpha$ changing from -0.46 to 0.08
\citealt{marrone:2006b}), upper limit for near-IR quiescent luminosity
(e.g. \citealt{genzel:2003}), an upper limit for X-ray luminosity $L_X <
2.4 \times 10^{33} [erg s^{-1}]$ ({\citealt{baganoff:2003}), size of the
image at 230 GHz ({\citealt{doeleman:2008}). We compare only
time-averaged SEDs and time-averaged images to the observations.
\section{Results}
\subsection*{Best-bet model}
We vary three (free) model parameters: spin of the black hole ($a_*$=0.5,
0.75, 0.84, 0.94, 0.96, 0.98), the ion to electron temperature ratio
($T_i/T_e=1, 3, 10$), and inclination of the observer with respect to
the spin axis ($i=5^{\deg}, 45^{\deg}$, and $85^{\deg}$). In
Table~\ref{tab1} we show a set of accretion flow models in which SEDs
are consistent with multiwavelength data as observed at at least one
inclinations. We find the ``best-bet'' model, that best satisfies all
the observational constraints (including VLBI 230 GHz size constrains)
has spin $a_*=0.94$, $T_i/T_e=3$ and $i=85^{\deg}$. All other models
are inconsistent with the observed SED or only marginally consistent
with the VLBI size measurement.
\begin{table}[!ht]
\smallskip
\begin{center}
{\small
\begin{tabular}{cccccccc ccc}
\tableline
run & $a_*$ & $i$ &$ < \dot{M}\cdot 10^{-9}>$ & $\alpha$ & $\log_{10} L_X$
&$\eta$&cons. \\
& & [deg] &$[M_{\odot}/yr^{-1}]$&& $[erg s^{-1}]$ &&w/
obs.?
\\
\tableline
& & 5 & 1.9 & -1.68& 31.5 & $3.5 \times 10^{-2}$& NO\\
D3 & 0.94 & 45& 1.7 & -1.27& 31.8 & $3.1 \times 10^{-2}$& NO\\
& & 85& 1.86 & -0.44& 32.9& $3.4 \times 10^{-2}$& YES\\
\\
& & 5 & 90.8 & -1.37& 30.1 &$5.4 \times 10^{-4}$& NO\\
A10 & 0.5 & 45& 117.1& -0.2 & 31.4 &$6.7 \times 10^{-4}$&YES\\
& & 85& 369.0& 1.38 & 33.5 &$1.6 \times 10^{-3}$& NO\\
\\
& & 5 & 38.3 & -1.05& 30.0 &$1.6 \times 10^{-3}$& NO\\
B10 & 0.75 & 45& 50.2 & 0.04 & 31.8 &$2.0 \times 10^{-3}$&YES\\
& & 85& 190.6& 1.49 & 34.6 &$6.9 \times 10^{-3}$& NO\\
\\
& & 5 & 19.5 & -1.15& 31.4 &$5.1 \times 10^{-3}$& NO\\
C10 & 0.875 & 45& 23.6 & -0.07& 32.0 &$6.2 \times 10^{-3}$&YES\\
& & 85& 41.4 & 1.19 & 34.2 &$1.7 \times 10^{-2}$& NO\\
\\
& & 5 & 13.7 & -0.93 & 31.8 &$1.1 \times 10^{-2}$& NO\\
D10 & 0.94 & 45& 15.2 & -0.05 & 32.3 &$1.1 \times 10^{-2}$&YES\\
& & 85& 31.2 & 1.17 & 34.4 &$2.5 \times 10^{-2}$& NO\\
\\
& & 5 & 13.6 & -0.40 & 32.5 &$2.6 \times 10^{-2}$& YES\\
E10 & 0.97 & 45& 14.3 & 0.2 & 33.1 &$2.8 \times 10^{-2}$&NO\\
& & 85& 26.5 & 1.19 & 35.2 &$5.1 \times 10^{-2}$& NO\\
\\
& & 5 & 9.07 & -0.6 & 32.7 &$5.3 \times 10^{-2}$& YES\\
F10 &0.98 & 45& 9.16 & -0.08 & 33.2 &$5.2 \times 10^{-2}$&NO\\
& & 85& 15.5 & 1.12 & 35.4 &$9.0 \times 10^{-2}$& NO\\
\tableline
\end{tabular}
}
\caption{Chosen set of models with temperature ratio 3 and 10 which SEDs are
consistent with observations at least at one observation angle. The columns from left
to right are: run ID (the number is the run name indicates $T_i/T_e$),
dimensionless spin of the black hole, inclination angle of the observer with
respect to the black hole spin axis, averaged rest mass accretion rate,
$\alpha$ spectral slope between 230-690 GHz ($F\sim\nu^{\alpha}$), and
luminosity in the X-rays (at $\nu \sim 10^{18}$ Hz), the radiative efficiency
$\eta=L_{\rm BOL} / \dot{M} c^2$, and whether the model is consistent with the
data. We do not show models with $T_i/T_e=1$ here because they are all
inconsistent with the observations.}\label{tab1}
\end{center}
\end{table}
In Figure~\ref{fig1} we show the SED of our best-bet model. The SED
peaks around 690 GHz due to thermal synchrotron emission. Below 100 GHz
it fails to fit the data because that emission is produced outside the
simulation volume. A second peak in the far UV is due to the first
Compton scattering order, at higher energies the photons are produced as
an effect of two or more scatterings.
\begin{figure}[!ht]
\begin{center}
\includegraphics[clip,scale = 0.3, angle = 0]{proc_fig2a.eps}
\includegraphics[clip,scale = 0.3, angle = -90, origin=c]{proc_fig2b.eps}
\end{center}
\caption{
Left panel: SEDs computed based on a single time slice $t=1680 G M/c^3$ (thick
line) (see Figure~\ref{fig2} for the distributions of the physical variables
corresponding to the same time) along with the time averaged spectrum (thin
line) of our best-bet model. Observational points are taken from:
\citealt{falcke:1998}, \citealt{an:2005}, \citealt{marrone:2006a} at radio,
\citealt{genzel:2003} at NIR (1.65, 2.16, and 3.76 ${\rm \mu m}$) and
\citealt{baganoff:2003} at X-rays (2-8 keV). The upper limits in the NIR band
are taken from \citealt{melia:2001} (30, 24.5 and 8.6 ${\rm \mu m}$),
\citealt{schoedel:2007} (8.6 $\mu m$) and \citealt{hornstein:2007} (2 ${\rm
\mu m}$). The points in the NIR at flaring state are from
\citealt{genzel:2003} (1.65, 2.16, and 3.76 ${\rm \mu m}$) , and
\citealt{dodds:2009} (3.8 ${\rm \mu m}$). An example of X-ray flare ($L_X = 1
\times 10^{35}$ $erg s^{-1}$) is taken from \citealt{baganoff:2001}. Right
panel: corresponding to SED in the left panel Image of the accretion flow at
$230 GHz$ correspindign to $t=1680 G M/c^3$.
Intensities are given in units of $erg s^{-1} {\rm
pixelsize^{-2} Hz^{-1} sr^{-1}}$, where the pixel size is $0.82 \mu as$. The
image shows inner $40 GM/c^2$. The white circle marks FWHM=37 $\mu as$ of a
symmetric Gaussian brightness profile centered at the image centroid.
}\label{fig1}
\end{figure}
In Figure~\ref{fig2}, we show the dynamical model corresponding to our
best-bet model. We show maps of number density density, magnetic field
strength and electron temperature ($\theta_e=k_b T_e/m_ec^2$) on a
single time slice.
\begin{figure}[!ht]
\begin{center}
\includegraphics[scale = 0.6, angle = -90]{proc_fig3.eps}
\end{center}
\caption{
Accretion flow structure in our best-bet model with $a_* = 0.94$ and with
$T_i/T_e=3$ (model D3). The number density overplotted with the velocity
field,
the magnetic field strength, and
the electron temperature are shown in the left, middle, and right panel
respectively. The axis units are $G M/c^2$.The figure shows a single
time slice.}\label{fig2}
\end{figure}
Figure~\ref{fig3} maps the points of origin for photons below and in the
synchrotron peak (100-690 GHz), in the NIR ($10^{13}-10^{14} Hz$), and in the
X-ray (2-8 keV). The figure again corresponds to a single time slice
from our best-bet model presented in Figures~\ref{fig1},~\ref{fig2}.
Most of the submillimeter emission originates near the mid plane at $4 <
r c^2/(G M) < 6$. NIR photons are produced in the hot regions close to
the innermost circular orbit ($r_{ISCO}\approx 2.04 G M/c^2$). Photons
in the X-ray band are produces mainly by scatterings in the hottest
parts of the disk also close to the $r_{ISCO}$. A small fraction of
photons are emitted from the funnel wall at large radii ($15-40
GM/c^2$).
\begin{figure}[!ht]
\begin{center}
\includegraphics[scale = 0.6, angle = -90]{proc_fig4.eps}
\end{center}
\caption{
Maps showing the point of origin for photons in our best-bet model. We show
the logarithm of the sum of the photon weights in each zone, which is
proportional to the number of photons seen at $100-690 GHz$ (left),
$10^{13}-10^{14} Hz$ (middle) and 2-8 keV (right) band. The axis scale units are
$GM/c^2$.
The figure presents a single time slice at $t=1680 G M/c^3$. Note that gray
scale bands differ in scales.}\label{fig3}
\end{figure}
\subsection*{Summary of parameter survey}
We can draw the following general conclusions from our work: (1) Very
few of the time averaged SEDs based on a single-temperature
($T_i/T_e=1$) models produce the correct spectral slope between 230-690
GHz. The exception is edge-on tori ($i=85^{\deg}$) around a fast
spinning black hole ($a_*=0.98,0.96$), but these models overproduce NIR
and X-ray flux.
(2) For $T_i/T_e=3$ the only one model with $a_*=0.94$ at $i=85^{\deg}$
is consistent with all observational constraints. For $i=85\deg$, models
with spins below $a_* = 0.94$ are ruled out by the inconsistent spectral
slope, and models with higher spins ($a_*>0.94$), although consistent
with the observed $\alpha$, overproduce the quiescent NIR and X-ray
emission. All models with $T_i/T_e=3$ observed at $i=5^{\deg}$ and
$45^{\deg}$ are ruled out by the inconsistent $\alpha$.
(3) For $T_i/T_e=10$, we find that all models with $i=85^{\deg}$ are
ruled out by both incorrect $\alpha$ and violation of NIR and X-ray
limits. For lower inclination angles ($i=5^{\deg},45^{\deg}$) a few
models (E10 and F10 with $i = 5\deg$, and A10, B10, C10, and D10 at $i =
45\deg$) reproduce the observed $\alpha$. These models are consistent
with X-rays and NIR limitations. Models E and F for $i=45\deg$ are
ruled out by NIR and X-ray limitations whereas models A10, B10, C10 and
D10 for $i=5\deg$ produce $\alpha$ which is too small.
(4) The dependence on $a_*$ arises largely because as $a_*$ increases
the inner edge of the disk --- the ISCO -- reaches deeper into the
gravitational potential of the black hole, where the temperature and
magnetic field strength are higher. In the disk mid-plane, the
temperature is proportional to the virial temperature and scales with
radius $\Theta_e \propto 1/r$. $B \propto 1/r$, while the density $\sim
r$, below the pressure maximum (longer simulations in 3D show a more
relaxed, declining density profile). Holding all else constant this
implies a higher peak frequency for synchrotron emission, a constant
Thomson depth (in our models the Thomson depth at the ISCO is roughly
constant, since the path length $1/\sim r_{ISCO}$ but the density $\sim
r_{ISCO}$), and a larger energy boost per scattering $A \approx 16
\Theta_e^2$, as can be seen in comparing models with different spin in
Figure 4. The X-ray flux therefore increases with $a_*$ because
$\Theta_e$ at the ISCO increases.
(5) The dependence on $T_i/T_e$ is mainly due to synchrotron
self-absorption, which is strongest at high inclination. For example,
because the $i = 85^{\deg}$, $T_i/T_e = 10$ model is optically thick at
$230 GHz$ the emission is produced in a synchrotron photosphere well
outside $r_{ISCO}$. The typical radius of the synchrotron photosphere
ranges between 15 $G M/c^2$ for low spin models ($a_*=0.5,0.75$) and 8
$G M/c^2$ for high spin models ($a_*> 0.75$). The $230 GHz$ flux can
then be produced only with large ${\mathcal M}$; as ${\mathcal M}$
increases the optically thin flux in the NIR increases due to increasing
density and field strength. The scattered spectrum also depends on
$T_i/T_e$ since the energy boost per scattering is $\sim 16\Theta_e^2
\propto 1/(T_i/T_e)^2$.
(6) The inclination dependence is a relativistic effect. ${\mathcal M}$
is nearly independent of $i$ (it varies by $\sim 10\%$, except for
$T_i/T_e = 10$, which due to optical depth effects has much larger
variation), so models with different inclination are nearly identical.
Nevertheless the X-ray flux varies dramatically with $i$, increasing by
almost 2 orders of magnitude from $i = 5\deg$ to $i = 85^{\deg}$. This
occurs because Compton scattered photons are beamed forward parallel to
the orbital motion of the disk gas. The variation of mm flux with $i$
is due to self-absorption. The mm flux reflects the temperature and
size of the synchrotron photosphere. At lower $i$ the visible
synchrotron photosphere is hotter than at high $i$.
\section{Future prospects}
We have made a number of approximations that will be removed in future
models of Sgr A*. We plan to (1) add cooling, (2) run 3D (rather than
axisymmetric) models, (3) model a larger range of radii, (4) include a
population of nonthermal electrons, (5) eliminate the fast-light
approximation by doing fully time-dependent radiative transfer.
\acknowledgements
This work was supported by the National Science Foundation under grants
AST 00-93091, PHY 02-05155, and AST 07-09246, by NASA grant NNX10AD03G,
through TeraGrid resources provided by NCSA and TACC, and by a Richard
and Margaret Romano Professorial scholarship, a Sony faculty fellowship,
and a University Scholar appointment to CFG.
|
\section{\label{sec:1} Introduction}
We assume that the cosmological background geometry filled with matter
or radiation is characterized by well known physical laws, like equation of
states and thermodynamics. It leads to a solid physical description for the
Universe in its very early stages. Such a description, especially in the Early
Universe, is favored, because we so far have no observational evidence against
it. Other components of the cosmological geometry, like dark matter and dark energy
wouldn't matter much during these early stages. Therefore, we can disregard them.
In this work, we introduce a toy model based on thermodynamical approaches to
describe the Early Universe. We disregard all phase transitions and assume
that the matter filling the background geometry was likely formed as free
gas. We apply the laws of thermodynamics and fundamentals of
classical physics to derive expressions for the basic cosmological quantities,
such as the Hubble parameter $H(t)$, the scaling factor $a(t)$ and the
curvature parameter $k$. We compare them with the Friedmann-Robertson-Walker
(FRW) model and Einstein field equations.
In this treatment, we apply the standard cosmological model and use natural
units in order to gain global evidences supporting the FRW model, although we
disregard the relativistic and microscopic effects. The various forms of matter
and radiation are homogeneously and isotropically distributed. We use non-relativistic
arguments to give expressions for the thermodynamic quantities in the Early Universe, which
obviously reproduce essential parts of FRW model. We assume that the
Universe was thermal equilibrium and therefore the interaction rates exceed
the Universe expansion rate, which was slowing down with the time $t$. Also, we
assume that the expansion was adiabatic, i.e., no entropy and heat change took place.
Finally, we take into consideration two forms of the cosmic background
matter. The first one is ideal gaseous fluid, which is characterized by lack
of interactions and constant internal energy. The second one is viscous fluid,
which is characterized by long range correlations and velocity gradient along
the scaling factor $a(t)$.
\section{\label{sec:2} Expansion Rate in Non-Viscous Cosmology}
We assume that all types of energies in the Early Universe are heat, $Q$. In
such a closed system, the total energy is conserved, i.e.,
\ba \label{eq:dq}
dQ &=& 0 = dU +pdV,
\ea
where $U$ is internal energy, $p$ is pressure and $V$ stands for the
volume. $V$ can be approximated as a cube with sides equal to the scaling
factor $a$, i.e., $V=a^3$ or as a sphere with radius equal to $a$, i.e., $V=(4\pi/3) a^3$. In both cases,
$V\propto a^3$. Apparently, Eq.~(\ref{eq:dq}) is the first law of
thermodynamics. The expansion of the Universe causes a change in energy
density $\rho=U/V$, i.e., surely decreasing, which can be given as $d\rho=dU/V-U
dV/V^2$. In comoving coordinates, $U$ is equivalent to the mass $m$ and
consequently to the energy. From Eq.~(\ref{eq:dq}), we get
\ba \label{eq:dro}
d\rho &=& -3 (\rho+p)\frac{da}{a}.
\ea
Dividing both sides by an infinitesimal time element $dt$ results in
\ba \label{eq:drot}
\dot \rho &=& -3 (\rho+p)\, H,
\ea
which is the equation of motion from FRW model, which strongly depends on the
thermodynamic quantities, $\rho$ and $p$, i.e., the equation of state
(EoS). One dot means first derivative with respect to the time
$t$. $H$ is the Hubble parameter which relates velocities with distances;
$H=\dot a/a$.
The radiation-dominated phase is characterized by $p=\rho/3$ and
therefore Eq.~(\ref{eq:drot}) leads to $\rho\propto a^{-4}$. In the
matter-dominated phase, $p<<\rho$ and therefore $\rho\propto a^{-3}$, i.e.,
$\rho\propto V^{-1}$. The energy density $\rho$ is a function of temperature
$T$. Then, we can re-phrase the proportionality in the radiation-dominated
phase as $\rho\propto T V^{-1}$.
Neglecting both cosmological constant $\Lambda$
and curvature parameter $k$, we simply get that $H^2\propto \rho$. Then, the scaling
factor in the radiation-dominanted phase $a\propto t^{1/2}$ and in
matter-dominanted phase $a\propto t^{2/3}$. The results are depicted in
Fig.~\ref{fig:1}.
\section{\label{sec:3} Expansion Rate in Bulk Viscous Cosmology}
Let us assume that one particle of mass $m$ is located at a distance $a$ from
some point in the Universe. Such a particle will have, in the radial
direction, kinetic energy $m\dot a^2/2$. In the opposite direction, it is
affected by a gravitational force due to its mass $m$ and the mass inside the
sphere $M=(4\pi/3)a^3 \rho$. Then the particles's {\it gravitational}
potential energy is $-GMm/a$, where $G$ is the Newtonian gravitational
constant. The total energy is
\ba \label{eq:tEnr1}
E&=&\frac{1}{2}m\dot a^2 - G\frac{Mm}{a},
\ea
which can be re-written as
\ba \label{eq:adot1}
\dot a^2 + k &=& \frac{8\pi}{3} G \rho a^2,
\ea
Last equation is, exactly, the Friedman's first equation with curvature
parameter $k=-2E/m$, which apparently refers to negative curvatures associated with
various geometrical forms depending on both total energy $E$ and mass $m$. In the
Friedman's solution, $k$ can be vanishing or $+1$ or $-1$, referring to flat
or positively or negatively curved Universe, respectively~\cite{dverno}. Our toy model
agrees well with the negatively curved Friedman's solution, especially
when the particle mass $m$ equals two times the total energy, i.e., $m=2E$.
According to recent heavy-ion collision experiments~\cite{rhic1} and lattice QCD
simulations~\cite{nakam}, the matter under extreme conditions (very high temperature
and/or pressure) seems not to be, as we used to assume, an {\it ideal} free
gas. It is likely fluid, i.e., strongly correlated matter with finite heat
conductivity and viscosity coefficients (bulk and shear)~\cite{finiteta1}. Therefore, it is
in demand to apply this assumption on the background geometry in Early Universe. The cosmic background should not necessarily be filled with an ideal free gas. In previous
works~\cite{taw1,taw2,taw3}, we introduced models, in which we included finite
viscosity coefficient. The analytical solution of such models is a non-trivial
one~\cite{taw1,taw2,taw3}. In the present work, we try to approach the viscous
cosmology using simple models, in which we just use classical approaches. As
we have seen, the classical approaches work perfectly in the non-viscous
fluid. It is in order now to check the influence of viscous fluid on the
cosmological evolution. The simplicity of these approaches doesn't sharpen the
validity of their results. Surely, it helps to come up with ideas on reality
of the viscous cosmology. \\
We now assume that the test particle is positioned in a viscous surrounding.
Then the total energy, Eq.~(\ref{eq:tEnr1}), gets an additional contribution from
the viscosity work, which apparently resists the Universe expansion,
\ba \label{eq:tEnr2}
E&=&\frac{1}{2}m\dot a^2 - G\frac{Mm}{a} - \eta a^3 \frac{\ddot a}{a},
\ea
where $\eta$ is the bulk viscosity coefficient. We assume that the expansion
of the Universe is isotropic, i.e., symmetric in all directions. Consequently,
the shear viscosity coefficient likely vanishes. Comparing
Eq.~(\ref{eq:tEnr2}) with the Friedmann's solution leads to another expression
for the curvature parameter,
\ba \label{eq:k1}
k &=& -\frac{2E}{m}-\frac{2 \eta}{m}\frac{\ddot a}{\dot a} a^3.
\ea
Comparing Eq.~(\ref{eq:k1}) with the three values of $k$ given in the FRW model ($k={+1,0,-1}$)~\cite{dverno}, results in three expressions for the expansion rate $\dot a$ in the bulk viscous cosmology. \\
When $k=+1$, then the expansion rate or velocity reads
\ba \label{eq:adot2}
\dot a &=& \left(\frac{2E+m}{\eta}\right)\frac{1}{a^2},
\ea
It is positive everywhere and inversely proportional to $a^2$. It doesn't
depend on the comoving time $t$, directly. Apparently, its $t$-dependence is embedded in
the $t$-dependency of $E$, $m$ and $\eta$. The scaling factor itself is
\ba \label{eq:aoft2}
a(t) &=& \left(3\frac{2E+m}{\eta}\right)^{1/3} t^{1/3}.
\ea
In Fig.~\ref{fig:1}, we compare this result with the non-viscous fluid as given
in section~\ref{sec:2}, i.e., $a(t)\propto t^{1/2}$ for radiation-dominated
phase and $a(t)\propto t^{2/3}$ for matter-dominated phase. For simplicity, we
assume that all proportionality coefficients are equal. It is clear that the
scaling factor in the viscous cosmology is the slowest one. This would refer
to the fact that the viscosity likely resists the Universe
expansion. Increasing $\eta$ shrinks or shortens $a(t)$,
Eq.~(\ref{eq:aoft2}). At very small $t$, the expansion of the bulk viscous
Universe is much rapid than the other two cases (non-viscous).
From Eqs.~({\ref{eq:adot2}) and (\ref{eq:aoft2}), the Hubble parameter reads
\ba \label{eq:H1}
H(t) &=& \frac{1}{3t}.
\ea
Apparently, $H$ doesn't depend on any of the thermodynamic quantities. It is
always positive and decays with increasing $t$. \\
\begin{figure}
\includegraphics[width=8cm,angle=-90]{aoft-comp1.eps}
\caption{\footnotesize Scaling factor $a$ as a function of comoving time $t$
is depicted for viscous fluid (solid), radiation-dominanted (long dashed) and
matter-dominanted (dotted) phases.}
\label{fig:1}
\end{figure}
When $k=-1$, the expansion rate or velocity takes the form
\ba \label{eq:adotn}
\dot a &=& \frac{2E-m}{\eta}\frac{1}{a^2}.
\ea
It is only positive, i.e., the Universe is only expanding as long as
$m<2E$. Otherwise, the expansion rate or velocity
decreases. Eq.~(\ref{eq:adotn}) sets the limit of the Universe
contraction. This limit is reached, when the mass $m$ exceeds twice the total
energy $E$. In Early Universe, $E>>m$ and, consequently,
the Universe started explosively, although $k$ could have a negative value. Much later, $E$ decreases
according to this expansion and the matter production, meanwhile the mass $m$ gains more
and more contributions. At a certain point, the expansion rate turns to the
backward direction. It is necessarily to mention here that, this toy model
takes into account the visible energy and matter components only. The
invisible components are not included in it.
The scaling factor also depends on the total energy $E$ and mass $m$,
\ba \label{eq:an}
a(t) &=& \left(3 \frac{2E-m}{\eta}\right)^{1/3} t^{1/3}.
\ea
$a$ is positive as long as $m<2E$. Otherwise it switches to negative
values. Its time dependence, $a\propto t^{1/3}$, looks like the previous case
at $k=+1$, Eq.~(\ref{eq:aoft2}).
From Eqs.~({\ref{eq:adotn}) and (\ref{eq:an}), the Hubble parameter reads
\ba \label{eq:H2}
H(t) &=& \frac{1}{3t}.
\ea
As in the previous case, $k=+1$, $H$ is always positive and doesn't depend on
any of the thermodynamic quantities. \\
When $k=0$, the expansion rate or velocity will be
\ba \label{eq:aoft0}
\dot a &=& \frac{1}{2} \frac{E}{\eta}\frac{1}{a^2}.
\ea
In flat Universe, $\dot a$ does not depend on $m$. It increases with
increasing the total energy $E$ and decreasing the viscosity coefficient
$\eta$. Also, the scaling rate,
\ba \label{eq:a0}
a &=& \left(\frac{3}{2} \frac{E}{\eta}\right)^{1/3} t^{1/3},
\ea
depends on $E$ and $\eta$, only. It doesn't depend on the mass $m$, i.e., the mass production doesn't affect the scale factor or the expansion. From Eqs.~(\ref{eq:aoft0}) and (\ref{eq:a0})
\ba \label{eq:H3}
H(t) &=& \frac{1}{3\, t}.
\ea
In Fig.~\ref{fig:2}, we depict $H(t)$ from this model and compare it with the
two cases when the background matter is a non-viscous gaseous fluid. The
latter is likely dominated by radiation or matter, where $H=1/2t$ and
$H=2/3t$, respectively. We notice that $H$ in the viscous cosmology is faster
than $H$ in the non-viscous cosmology. Therefore, we conclude that the bulk viscosity causes
slowing down the Universe expansion. \\
\begin{figure}
\includegraphics[width=8cm,angle=-90]{Hoft-comp1.eps}
\caption{\footnotesize The Hubble parameter $H$ as a function of comoving time
$t$ is depicted for viscous fluid (solid) and ideal gas. The radiation- and
matter-dominanted ideal gasses are drawn as long-dashed and dotted lines,
respectively.}
\label{fig:2}
\end{figure}
At $k=0$, the total energy is likely dominated by the viscosity work
\ba
E &=& -\eta a^3 \frac{\ddot a}{\dot a}.
\ea
Plugging this equation into the scaling factor, Eq.~(\ref{eq:a0}), leads to deceleration
\ba
\ddot a &=& -\frac{2}{3} \frac{\dot a}{t}.
\ea
Two dots refer to second derivative with respect to time $t$. \\
So far, we conclude that filling the cosmic background geometry with bulk
viscous fluid strongly affects (moderates) the expansion rate of the
Universe. The evolution of the scaling factor $a$ is damped, when $\eta$
increases. The curvature parameter $k$, which appears as a constant in FRW
model and Einstein equations~\cite{dverno}, depends on the total energy $E$,
the particle
mass $m$ and the viscosity coefficient $\eta$. In the positively curved
Universe, $a$ increases with increasing $E$ and $m$. In the negatively
curved Universe, $a$ increases only as long as $2E>m$. Otherwise, it
decreases causing Universe contraction. We conclude that the scaling factor in
the flat Universe depends on $E/\eta$, but not on the mass constent, $m$. \\
\section{\label{sec:3b} Energy Density in Bulk Viscous Cosmology}
Including the work of bulk viscosity into the first law of thermodynamics, Eq.~(\ref{eq:dq}), results in
\ba \label{eg:du-eta}
dU &=& -\left(p\, dV + \eta\, a^3\, \frac{\ddot a}{\dot a}\, dV\right).
\ea
Apparently, the evolution of energy density depends on the Hubble parameter $H$ and the viscosity coefficient $\eta$,
\ba \label{eq:rhovis1}
\dot \rho &=& -\left(3(p+\rho) + 3\, \eta\, a^3\, \frac{d \dot a}{d a} \right)\, H.
\ea
Comparing this evolution equation with the one in the Eckart~\cite{Eck40}
relativistic cosmic fluid leads to a direct estimation for the bulk viscous
stress $\Pi$. The conservation of total energy density requires that the bulk
viscous stress equals to the work of bulk viscosity, i.e., $\Pi=\eta a^3 d\dot
a/d a$.
In the radiation-dominanted phase, EoS reads $p=\rho/3$ and
Eq.~(\ref{eq:rhovis1}) can be solved in comoving time $t$ by utilizing our
previous result on $H(t)$, Eq.~(\ref{eq:H3}) for instance,
\ba \label{eg:rho-eta3}
\rho &=& -\ln t^{(12p-\eta/3)}.
\ea
On one hand, it implies that $\rho$ diverges at $t=0$. On the other hand, the
viscosity coefficient $\eta$ seems to moderate the evolution of the total
energy density. Obviously, this result strongly depends on
EoS~\cite{taw1,taw2,taw3}, which is different in the different phases of Early
Universe, i.e. differs with $t$.
Eq.~(\ref{eq:rhovis1}) is consistent with the second law of thermodynamics at
non-negative entropy production, $S^i_{;i}=\Pi^2/\eta T \ge0$. In this model,
$S^i_{;i} \propto V H/T$.
The Friedmann's second solution in flat Universe, whose background geometry is
filled with a non-viscous fluid,
\ba \label{eq:2nd1}
\frac{\ddot a}{a} &=& - \frac{4 \pi}{3} \, G \, (p+\rho).
\ea
seems to follow the second law of thermodynamics. To show this, let us take
the time derivative of last expression. Then $d (\rho\, a^3)\equiv - a^2
(p+\rho) d a$. Depending on EoS, for instance in de Sitter Universe, last
equivalence can be re-written as
\ba
d (\rho\, a^3) &=& - 3 a^2 p\, da.
\ea
It is nothing but the second law of thermodynamics ($dU=-pdV+TdS$) of an adiabatic system, i.e., the expansion is thermally reversible and obviously doesn't affect the entropy content, $dS=0$.
\section{\label{sec:4} Hubble Parameter in Quantum Cosmology}
Let us suppose that $N$ particles are adhered to a cubic or spherical volume,
i.e., $V\propto a^3$. The particles are distributed according an occupation
function, which depends on their quantum numbers and correlations. According
to the standard cosmological model, they are allowed to expand in a
homogeneous and isotropic way. We suppose that particles have no
interactions. Then, the energy of a single particle $E=(p^2+m^2)^{1/2}$, where
momentum $\vec{p}=\frac{h}{a}(n_1\hat{x}+n_2\hat{y}+n_3\hat{z})$. In natural
units, $h=2\pi$. The state density in momentum space
$a^3/h^3=V/(2\pi)^3$. From the integral of particle density in phase space, we
get the particle density in ordinary space and, therefore,
\ba
a^3 &=& N \left.\frac{2\pi^2}{g} \right/\left.\int_0^{\infty} \frac{p^2
dp}{e^{\frac{E-\mu}{T}}\pm 1} \right.,
\ea
where $\mu$ is the chemical potential and $g$ is the degeneracy factor.
Taking the time derivative (equivalent to $1/T$) and dividing both sides by the scaling factor $a$ results in
\ba
H(T,t) &=& \frac{1}{n\,T^2} \left\{\frac{1}{6} \frac{g}{2\pi^2}
\int_0^{\infty} \frac{\mu-E}{1+ \cosh(\frac{E-\mu}{T})} \;p^2\,dp\right\}
\frac{dT}{dt}, \label{eq:Hquant2}
\ea
where $n$ is the particle number density. $n$ depends on the intensive state
variables, $T$, $p$ and $\mu$. It implies that $H$ depends on $1/n T^2$ and
the time derivative, $dT/dt$, besides the integral, which can be calculated,
numerically, in dependence on $T$ and $\mu$. It is obvious that the expansion
of the Universe is driven by generating new states. \\
When assuming that the background geometry is filled with a relativistic
Boltzmann's gas, then the equilibrium pressure and energy density at vanishing
viscosity are given as
\ba \label{eq:pE1}
p(m,T) &\approx&nT, \nonumber \\
\epsilon(m,T) &\approx& n \left(\frac{3T}{m}+ \frac{K_1(m/T)}{K_2(m/T)}
\right)\, m,
\ea
where $K_i$ is the $i$-th order modified Bessel function. At equilibrium,
the entropy is maximum. At vanishing chemical potential, the Hubble
parameter in Eq.~(\ref{eq:Hquant2}) reads
{\footnotesize\bf
\ba \label{eq:Hquant3}
H &=& -\frac{g}{2\pi^2}\frac{1}{6n} \left[p E + 4T {\cal M}
\tanh^{-1}\left(\frac{2 p T}{E {\cal M}}\right) -
\left(m^2+8T^2\right)\ln(2p+2E)
\right] \frac{dT}{dt}, \nonumber
\ea }
where ${\cal M}=(m^2+8T^2)^{1/2}$. In the relativistic limit, i.e.,
$m\rightarrow 0$,
\ba \label{eq:Hquant4}
H &=& -\frac{g}{2\pi^2}\frac{1}{6n} \left[p^2 + 8 T^2\left(\sqrt{2}
\tanh^{-1}\frac{1}{\sqrt{2}} - \ln(4p)\right)
\right] \frac{dT}{dt}.
\ea
We conclude that $H$ in non-viscous quantum cosmology $H$ depends on the
intensive state quantity $T$, its decay with the time $t$, state density in
momentum space and both of momentum and number of occupied
states~\cite{taw-new}.
\section{\label{sec:5} Bulk Viscosity in Quantum System}
In this section, we give estimates for the bulk viscous coefficient in both of
quark-gluon plasma and hadrons, which can be inserted in Eq.~(\ref{eg:du-eta})
to calculate the expressions given in
Eqs.~(\ref{eq:rhovis1})~and~(\ref{eg:rho-eta3}).
In the relativistic Boltzmann limit, the bulk viscosity~\cite{ilg1, deGroot}
reads
\ba \label{eq:etaa1}
\eta\left(\frac{m}{T}\right) &=& \frac{m^2}{96 \pi^2 \sigma} \;
\frac{\left\{K_2\left(\frac{m}{T}\right)\left[(5-3\gamma)\hat{h}-3\gamma
\right] \right\}^2} {2T\,
K_2\left(\frac{2m}{T}\right)+m\,K_3\left(\frac{2m}{T}\right)},
\ea
which implies that $\eta$ doesn't depend on the extensive state quantity $n$,
which gives the number of occupied states in momentum space. $\eta$ depends on
the ratio of heat capacities and enthalpy per particle, which are given by the
auxiliary functions $\gamma/(\gamma-1)=5\hat{h}-\hat{h}^2+T^2/m^2$ and
$\hat{h}=T[K_3(m/T)/K_2(m/T)]/m$, respectively. Also, it depends on mass $m$,
temperature $T$ and the cross section $\sigma$. The latter has been given in
Ref.~\cite{hisc1} and is assumed to be constant for all states or particles.
At $m=5$ and constant $\sigma$, Eq.~(\ref{eq:etaa1}) is drawn in
Fig.~\ref{fig:4}. It is clear that $\eta(m/T)$ has two singularities, one at
$T=0$ and another one at $m/2$. It has a minimum value, at a temperature
slightly below $m/2$. At much higher temperatures, $\eta(m/T)$ increases
linearly with increasing $T$. In the high-$T$ region, Eq.~(\ref{eq:etaa1}) is
likely no longer valid. \\
\begin{figure}
\begin{center}
\includegraphics[width=8cm]{etaa1.eps}
\caption{\footnotesize At constant particle mass $m$ and cross section
$\sigma$, $\eta$ is drawn againest $T$. There are two singularities at $T=0$
and at $m/2$. $\eta(m/T)$ is minimum at temperatures slightly below $m/2$. }
\label{fig:4}
\end{center}
\end{figure}
In the Hagedorn picture, the particles, at very high energies, can be treated
non-relativistically. In this limit, we express the partition function in mass
spectrum $\rho(m)$.
\ba
\ln Z &=& \frac{g}{2\pi^2}\,V\,T^3 \sum_{n=1}^{\infty} \rho_n(m) \frac{1}{n^2}
\left(\frac{m}{T}\right)^2 K_2\left(\frac{n\,m}{T}\right).
\ea
The mass spectrum $\rho(m) dm$ give the number of states between $m$ and
$m+dm$. The non-relativistic part of the energy-momentum tensor is~\cite{ostr}
\ba
T^{i j} &=&\frac{\tau}{(2\pi)^3} \left(\frac{i_k}{T}\right)_{,l}\int
\frac{p^{i}\, p^{j}\, p^{k}\, p^{l}}{m E} e^{-E/T} \,\rho(m)\, dm\; d^3p,
\ea
where $\tau$ is the relaxation time.
Taking into account the asymptotic behavior allows us to derive the bulk
viscous term in $T^{i j}$,
\ba \label{eq:etacs}
\eta &=& \frac{5}{3}{\cal A} \frac{\tau T^{5/2}}{(2\pi)^{3/2}}
\ln\left(c_s^{-2}\right),
\ea
where ${\cal A}$ is constant and $c_s=(\partial p/\partial \epsilon)^{1/2}$ is
the speed of sound, which characterizes the propagating of signals in the
cosmological background matter of the Early Universe. In the relativistic
limit, the partial derivatives in $c_s$ are taken, adiabatically, i.e., at
constant heat (or energy as we assumed in this model).
$\ln(c_s^{-2})$ in Eq,~(\ref{eq:etacs}) can roughly be estimated, when we
approximate the thermodynamic quantities $p$ and $\epsilon$,
Eq.~(\ref{eq:pE1}). We assume that $p$ and $\epsilon$ are not changing with
the bulk viscous coefficient $\eta$, then
\ba
c_s^{-2}\left(\frac{m}{T}\right) &=&
\frac{\left(\frac{m}{T}\right)^2}{2 K_2^2 \left(\frac{m}{T}\right)}
\left\{ K_0^2 \left(\frac{m}{T}\right) - 2
\left(\frac{m}{T}\right)^{-1} K_0^2 \left(\frac{m}{T}\right)
K_1^2 \left(\frac{m}{T}\right) - \right. \nonumber \\
& & \left. 2 \left[1+4\left(\frac{m}{T}\right)^{-2}\right]
K_1^2\left(\frac{m}{T}\right) + \left[1+6\left(\frac{m}{T}\right)^{-2}\right]
K_2^2\left(\frac{m}{T}\right) \right\}. \nonumber
\ea
Using the dimensionless ratio $m/T$, last equation can be calculated,
numerically. This is illustrated in Fig.~\ref{fig:3}. The asymptotic value,
$\ln(3)$, is fulfilled at high $T$. When $T\rightarrow m$, the function drops
to a minimum value. It diverges as long as $T<m$. The inverse of the
relaxation time gives the drag coefficient of the background matter. In
relativistic limit, we can model the relaxation time. For strongly coupled
${\cal N}=4$ SYM~\cite{mald}, $\tau=f(1/T)$,
\ba\label{eq:tau1}
\tau &=& \frac{2-\ln(2)}{2\pi T}.
\ea
\begin{figure}
\begin{center}
\includegraphics[width=8cm]{lncs2_2.eps}
\caption{\footnotesize At a constant particle mass, $\ln\left(c_s^{-2}\right)$
is depicted as a function of $T$. Decreasing $T$ lowers $\ln(c_s^{-2})$
below its asymptotic limit, $\ln(3)$. When $T\rightarrow m$,
$\ln\left(c_s^{-2}(m/T)\right)$ drops to a minimum value. }
\label{fig:3}
\end{center}
\end{figure}
To keep fitting scope of present work, we leave for a future
work~\cite{taw-new}, the numerical estimates of Hubble paramter in the quantum
cosmology with (section~\ref{sec:4}) and without (section~\ref{sec:5}) bulk
viscousity.
\section{\label{sec:6} Conclusions}
We have shown that the Universe, which is characterized by the FRW model,
apparently obeys the laws of thermodynamics. We used classical assumptions in
order to derive the essential cosmological parameters, Hubble parameter $H$,
scaling factor $a$ and curvature constant $k$. In doing so, we have assumed
that the background matter is filled with an ideal thermal gas. Such a matter
is homogeneously and isotropically distributed inside the available
cosmological geometry. For simplicity, we assume that no interactions or phase
transitions took place.
The first gaol of this work is to study the effects of including finite bulk
viscosity on the Early Universe using classical approaches. We started from
the same assumptions as we did with the ideal thermal gas. We found
considerable changes in all cosmological parameters. Comparing our results
with the FRW model results in that the time-dependence of Hubble parameter and
scaling factor is slower than that of taking the background matter as an ideal
thermal gas.
Also, we have found a strong dependence of $k$ on the thermodynamic
quantities, like total energy $E$ and bulk viscosity $\eta$. The relation
between $E$ and the mass $m$ determines the time scale, at which the negative
curvature sets on. The total energy in the flat Universe is characterized by a
dominant work of the bulk viscosity. When this takes place, the Universe
decelerates. Otherwise, the expansion is positive. The expansion rate is
directly proportional to $E$ and $\eta$. As for thermodynamics of the Early
Universe, we have found that the time evolution is affected by the bulk
viscosity coefficient. Should this model be considered acceptable, essential
modifications in the astrophysical observations are expected.
The second goal of this work is to check the cosmological parameters in a
quantum system. We started from basic assumptions of quantum mechanics and
statistical physics. We expressed the Hubble parameter (and scaling factor in
a straightforward way) in dependence on the infinitesimal changes in both
phase and momentum spaces. We found that the Hubble parameter depends on $1/T$
and the time derivative of $T$. Based on this toy model, it is clear that the
expansion of the Universe is derived by the generation of new states. In the
relativistic limit, Hubble parameter depends on momentum space and the number
of occupied state, besides the decay of $T$.
Finally, we have studied the bulk viscosity in low-$T$ and high-$T$
regimes. For the first regime, the bulk viscosity diverges at vanishing $T$
and at $T\approx m/2$, where $m$ is mass. For the high-$T$ regime, the bulk
viscosity decreases with increasing $T$. Its asymptotic value is reached, when
the speed of sound approaches its asymptotic limit. At vanishing $T$, the
speed of sound diverges.
\section*{Acknowledgments}
This work is based on an invited talk given at the ``Second IAGA-Symposium'',
which has been held in Cairo-Egypt from 4th till 8th January 2010. I like to
thank the organizers, especially Professor A. A. Hady for the kind invitation.
|
\section{Introduction}
It is well known that the $S^3$-bundles over $S^2$ are classified by $\pi_1(SO(4))={\mathbb Z}_2$. So there are exactly two such bundles, the trivial bundle $S^2\times S^3$ and one non-trivial bundle $X_\infty$ (in Barden's notation \cite{Bar65}). They are distinquished by their second Stiefel-Whitney class $w_2\in H^2(M,{\mathbb Z}_2)$.
There are infinitely many Sasaki-Einstein metrics\footnote{For basic material concerning Sasakian geometry I refer to our recent book \cite{BG05}.} on $S^2\times S^3$ which belong to a toric contact structure \cite{GMSW04a,GMSW04b,CLPP05,MaSp05b}. These, of course, are all extremal, but they have $c_1({\mathcal D})=0$ where ${\mathcal D}$ is the contact bundle. There is a more or less obvious constant scalar curvature extremal Sasakian metric on $S^2\times S^3$ which is not Sasaki-Einstein and has first Chern class $c_1({\mathcal D})=2(k_1-k_2)\alpha$ for every pair $(k_1,k_2)$ of relatively prime positive integers. Here $\alpha$ is a generator of $H^2(S^2\times S^3,{\mathbb Z})$, and without loss of generality we can assume that $k_1> k_2$. We recover the well known Kobayashi-Tanno homogenous Einstein metric in the case $(k_1,k_2)=(1,1)$. The Sasakian structures are constructed from the K\"ahler form $\omega_{k_1,k_2}=k_1\omega_1+k_2\omega_2$ on $S^2\times S^2$ with the product complex structure, where $\omega_1(\omega_2)$ are the standard symplectic forms on the first (second) factor. The metric corresponding to this K\"ahler form has constant scalar curvature. One then forms the $S^1$-bundle over $S^2\times S^2$ whose cohomology class is $[\omega_{k_1,k_2}]$. The constant scalar curvature K\"ahler metrics lifts to a constant scalar curvature Sasakian metric on $S^2\times S^3$ which is homogeneous, hence toric. So there is a three dimensional Sasaki cone $\kappa$ as described in \cite{BGS07b}, and by the openness theorem of \cite{BGS06} there is an open set of extremal Sasakian metrics in $\kappa$. I want to emphasize that although extremal quasi-regular Sasakian metrics are always lifts of extremal K\"ahlerian orbifold metrics, it is NOT true that the openness theorem for extremal Sasaki metrics is obtained by simply lifting the openness theorem \cite{LeSi93b,LeSi94} for extremal K\"ahler metrics. In this note I shall prove that there are many other extremal Sasakian metrics on $S^2\times S^3$ belonging to the same contact structure.
In contrast to the situation of $S^2\times S^3$, until now there are no known extremal Sasakian metrics on $X_\infty$. Here I also show that $X_\infty$ admits many extremal Sasakian structures belonging to the same contact structure.
\section{Review of Extremal Sasakian Metrics}
Recall \cite{BG05} that a Sasakian structure ${\oldmathcal S}=(\xi,\eta,\Phi,g)$ on a smooth manifold $M$ is a contact metric structure such that $\xi$ preserves the underlying almost CR structure $({\mathcal D},J)$ defined by ${\mathcal D}=\ker \eta$ and $J=\Phi|_{\mathcal D}$, and the almost CR structure is integrable. Now we deform the contact 1-form by $\eta\mapsto \eta(t)=\eta+t\zeta$ where $\zeta$ is a basic 1-form with respect to the characteristic foliation ${\mathcal F}_\xi$ defined by the Reeb vector field $\xi.$ Here $t$ lies in a suitable interval containing $0$ and such that $\eta(t)\wedge d\eta(t)\neq 0$. This gives rise to a family of Sasakian structures ${\oldmathcal S}(t)=(\xi,\eta(t),\Phi(t),g(t))$ that we denote by ${\mathcal S}(\xi, \bar{J})$ where $\bar{J}$ is the induced complex structure on the quotient bundle $\nu({\mathcal F}_\xi)=TM/L_\xi$ by the trivial line bundle generated by $\xi.$ As the notation suggests we always assume that ${\oldmathcal S}(0)=(\xi,\eta(0),\Phi(0),g(0))={\oldmathcal S}$.
Note that this deforms the contact structure (hence, the CR structure) from ${\mathcal D}$ to ${\mathcal D}_t=\ker \eta(t)$, but they are isomorphic as complex vector bundles and isotopic as contact structures by Gray's theorem (cf. \cite{BG05}). In fact, each choice of 1-form defines a splitting
$TM=L_\xi+{\mathcal D}$ and an isomorphism ${\mathcal D}\approx \nu({\mathcal F}_\xi)$ of complex vector bundles. The complex structure $J$ on ${\mathcal D}$ defines a further splitting of the complexified bundle ${\mathcal D}\otimes {\mathbb C}={\mathcal D}^{1,0}+{\mathcal D}^{0,1}$, and the usual Dolbeault type complexes with transverse Hodge theory holds \cite{BG05}, and the same for ${\mathcal D}_t$.
We assume that $M$ is compact of dimension $2n+1$ with a Sasakian structure ${\oldmathcal S}$, and note that the associated Riemannian metric is uniquely determined by $\eta$ and $\Phi$ as
$$g=d\eta\circ (\Phi\otimes {\mathchoice {\rm 1\mskip-4mu l} {\rm 1\mskip-4mu l) \oplus \eta\otimes \eta.$$
Following \cite{BGS06} we let $s_g$ denote the scalar curvature of $g$ and define
the ``energy functional'' $E:{\mathcal S}(\xi,\bar{J})\ra{1.4} {\mathbb R}$ by
\begin{equation}\label{var}
E(g) ={\displaystyle \int _M s_g ^2 d{\mu}_g ,}\,
\end{equation}
i.e. the $L^2$-norm of the scalar curvature. Critical points $g$ of this functional are called {\it extremal Sasakian metrics}. In this case we also say that the Sasakian structure ${\oldmathcal S}$ is extremal. Similar to the K\"ahlerian case, the Euler-Lagrange equations for this functional give \cite{BGS06}
\begin{theorem}\label{ELeqn}
A Sasakian structure ${\oldmathcal S}\in {\mathcal S}(\xi,\bar{J})$ is a critical point for the energy functional (\ref{var}) if and only if the gradient vector field $\partial^\#_gs_g$ is transversely holomorphic. In particular, Sasakian metrics with constant scalar curvature are extremal.
\end{theorem}
Here $\partial^\#_g$ is the $(1,0)$-gradient vector field defined by $g(\partial^\#_g\varphi,\cdot)= \bar{\partial}\varphi$.
It is important to note that a Sasakian metric $g$ is extremal if and only if the `transverse metric' $g^T=d\eta\circ (\Phi\otimes {\mathchoice {\rm 1\mskip-4mu l} {\rm 1\mskip-4mu l)$ is extremal in the K\"ahlerian sense which follows from the well known relation between scalar curvatures $s_g=s_g^T-2n$ where $s_g^T$ is the scalar curvature of the transverse metric. It follows that
\begin{proposition}\label{SasKah}
Let ${\oldmathcal S}=(\xi,\eta,\Phi,g)$ be a Sasakian structure on $M$ of dimension $2n+1$ and let $U\subset M$ be an open set such that $f:U\ra{1.5} {\mathbb C}^n$ is a local submersion. Then the restriction $g|_U$ is an extremal Sasakian metric if and only if $g^T$ viewed as a K\"ahlerian metric on $f(U)$ is extremal. In particular, if the Sasakian structure ${\oldmathcal S}$ is quasi-regular and $(\omega,\hat{J},h)$ is the induced orbifold K\"ahlerian structure on the quotient ${\oldmathcal Z}=M/{\mathcal F}_\xi$, then $g$ is Sasakian extremal if and only if $h$ is K\"ahlerian extremal. Moreover, $g$ has constant scalar curvature if and only if $h$ has constant scalar curvature.
\end{proposition}
Given a strictly pseudoconvex CR structure $({\mathcal D},J)$ of Sasaki type on a smooth manifold $M$ of dimenision $2n+1$, we consider the set ${\mathcal S}({\mathcal D},J)$ of Sasakian structures whose underlying CR structure is $({\mathcal D},J)$. The group ${\mathfrak C}{\mathfrak R}({\mathcal D},J)$ of CR transformation acts on ${\mathcal S}({\mathcal D},J)$, and the quotient space $\kappa({\mathcal D},J)$ is called the Sasaki cone \cite{BGS06} whose dimension satisfies $1\leq \dim \kappa({\mathcal D},J)\leq n+1$. For a strictly pseudoconvex CR structure $({\mathcal D},J)$ on a compact manifold the group ${\mathfrak C}{\mathfrak R}({\mathcal D},J)$ is compact except in the case of standard CR structure on the sphere $S^{2n+1}$ \cite{Sch95,Lee96} in which case it is isomorphic to $SU(n+1,1)$ \cite{Web77}. Thus, ${\mathfrak C}{\mathfrak R}({\mathcal D},J)$ has a unique maximal torus up to conjugacy. So, as discussed in \cite{Boy10a} it is often convenient to consider the `unreduced' Sasaki cone ${\mathfrak t}^+={\mathfrak t}^+({\mathcal D},J)$ where ${\mathfrak t}$ is the Lie algebra of a maximal torus in the group ${\mathfrak C}{\mathfrak R}({\mathcal D},J)$ of CR transformations. This is defined by
\begin{equation}\label{uSascone}
{\mathfrak t}^+=\{\xi'\in {\mathfrak t}~|~\eta(\xi')>0\}
\end{equation}
where $\eta$ is any 1-form representing ${\mathcal D}$, and is an open convex cone in ${\mathfrak t}$. It is related to the Sasaki cone $\kappa({\mathcal D},J)$ by $\kappa({\mathcal D},J)={\mathfrak t}^+({\mathcal D},J)/{\mathcal W}$ where ${\mathcal W}$ is the Weyl group of ${\mathfrak C}{\mathfrak R}({\mathcal D},J)$. By abuse of terminology I also refer to ${\mathfrak t}^+$ as the Sasaki cone. Note that with $({\mathcal D},J)$ fixed, choosing a Reeb field uniquely chooses a 1-form $\eta$ such that $\ker\eta={\mathcal D}$, and with $J$ also fixed $\Phi$, hence, $g$ are uniquely specified. Thus, we can think of the Sasaki cone ${\mathfrak t}^+({\mathcal D},J)$ as consisting of Sasakian structures ${\oldmathcal S}=(\xi,\eta,\Phi,g)$, so again by abuse of notation we can write ${\oldmathcal S}\in {\mathfrak t}^+({\mathcal D},J)$.
Let $\eta(t)=\eta+t\zeta$ where $\zeta$ is a basic 1-form that is invariant under the full torus $T$. Such a 1-form can always be obtained by averaging over $T$. So the Lie algebra ${\mathfrak t}$ does not change under such a deformation, but generally the Sasaki cone ${\mathfrak t}^+$ associated with ${\mathcal D}_t$ shifts, and we denote it by ${\mathfrak t}^+(t)$. We let ${\mathcal S}^T(\xi,\bar{J})$ denote the subset of ${\mathcal S}(\xi,\bar{J})$ consisting of $T$-invariant Sasakian structures. If ${\oldmathcal S}=(\xi,\eta,\Phi,g)$ is $T$-invariant, then ${\mathcal S}^T(\xi,\bar{J})$ consists of all deformations obtained by $\eta\mapsto \eta(t)=\eta+t\zeta$ with $\zeta$ invariant under $T$. We are interested in the case when the torus $T$ has maximal dimension and the Reeb vector field is an element of the Lie algebra ${\mathfrak t}$, that is, a toric contact structure ${\mathcal D}$ of {\it Reeb type}.
\begin{lemma}\label{invlem}
Let $(M,{\mathcal D},T)$ be a toric contact structure of Reeb type with Reeb vector field $\xi.$ Suppose also that there is an extremal representative ${\oldmathcal S}(t)=(\xi,\eta(t),\Phi_t,g_t)\in {\mathcal S}(\xi,\bar{J})$. Then ${\oldmathcal S}(t)\in {\mathcal S}^T(\xi,\bar{J})$.
\end{lemma}
\begin{proof}
Since the contact structure ${\mathcal D}$ is toric of Reeb type, there is a compatible $T$-invariant Sasakian structure ${\oldmathcal S}=(\xi,\eta,\Phi,g)$ by \cite{BG00b}, and suppose that ${\mathcal S}(\xi,\bar{J})$ has an extremal representative ${\oldmathcal S}(t)=(\xi,\eta(t),\Phi_t,g_t)$. By the Rukimbira Approximation Theorem, if necessary, there is a quasi-regular $T$-invariant Sasakian structure invariant close to ${\oldmathcal S}$, so we can assume that ${\oldmathcal S}$ is quasi-regular. But then by Proposition \ref{SasKah} the Sasakian deformation corresponds to a K\"ahlerian deformation on the base K\"ahler orbifold $M/{\mathcal F}_\xi$. By a theorem of Calabi \cite{Cal85} this deformed K\"ahler structure has maximal symmetry, and one easily sees that the corresponding deformed Sasakian structure ${\oldmathcal S}(t)$ also has maximal symmetry which implies that ${\oldmathcal S}(t)\in {\mathcal S}^T(\xi,\bar{J})$.
\end{proof}
Now let $({\mathcal D},J)$ be a strictly pseudoconvex CR structure of Sasaki type. We say that ${\oldmathcal S}\in {\mathfrak t}^+({\mathcal D},J)$ is an {\it extremal element}\footnote{In \cite{BGS06} this was called a canonical element, but I prefer to call it an extremal element.} of ${\mathfrak t}^+({\mathcal D},J)$ if there exists an extremal Sasakian structure in ${\mathcal S}(\xi,\bar{J})$. We also say that ${\oldmathcal S}$ is an extremal element of $\kappa({\mathcal D},J)$. So we can define the {\it extremal set} ${\mathfrak e}({\mathcal D},J)\subset \kappa({\mathcal D},J)$ as the subset consisting of those elements that have extremal representatives (similarly for ${\mathfrak t}^+({\mathcal D},J)$).
Recall the {\it transverse homothety} (cf. \cite{BG05}) taking a Sasakian structure ${\oldmathcal S}=(\xi,\eta,\Phi,g)$ to the Sasakian structure $${\oldmathcal S}_a=(a^{-1}\xi,a\eta,\Phi,ag+(a^2-a)\eta\otimes \eta)$$
for any $a\in {\mathbb R}^+$. It is easy to see that
\begin{lemma}\label{transhomoext}
Let ${\oldmathcal S}$ be a Sasakian structure. If ${\oldmathcal S}$ is extremal, so is ${\oldmathcal S}_a$, and if ${\oldmathcal S}$ has constant scalar curvature so does ${\oldmathcal S}_a$. In particular, if the extremal set ${\mathfrak e}({\mathcal D},J)$ is non-empty it contains an extremal ray of Sasakian structures.
\end{lemma}
A main result of \cite{BGS06} says that ${\mathfrak e}({\mathcal D},J)$ is an open subset of $\kappa({\mathcal D},J)$. Lemma \ref{transhomoext} says that ${\mathfrak e}({\mathcal D},J)$ is conical in the sense that it is a union of open cones, so it is not necessarily connected.
\section{Bouquets of Sasaki Cones and Extremal Bouquets}
There may be many Sasaki cones associated to a given contact structure ${\mathcal D}$ of Sasaki type. They are distinguished by their complex structures $J.$ As shown in \cite{Boy10a} to a compatible almost complex structure $J$ on a compact contact manifold $(M,{\mathcal D})$ one can associate a conjugacy class ${\mathcal C}_T({\mathcal D})$ of maximal tori in the contactomorphism group ${\mathfrak C}{\mathfrak o}{\mathfrak n}(M,{\mathcal D})$. Furthermore, almost complex structures that are equivalent under a contactomorphism give the same conjugacy class ${\mathcal C}_T({\mathcal D})$. So given inequivalent complex structures $J_l$ labelled by positive integers, we can associate unreduced Sasaki cones ${\mathfrak t}^+({\mathcal D},J_l)$, or the full Sasaki cones $\kappa({\mathcal D},J_l)$. This leads to
\begin{definition}\label{Sasbou}
We define a {\bf bouquet of Sasaki cones} ${\mathfrak B}({\mathcal D})$ as the union
$${\mathfrak B}({\mathcal D})=\cup_{l\in {\mathcal A}}\kappa({\mathcal D},J_l)$$
where ${\mathcal A}\subset {\mathbb Z}^+$ is an ordered subset. We say the it is an {\bf $N$-bouquet} if the cardinality of ${\mathcal A}$ is $N$ and denote it by ${\mathfrak B}_N({\mathcal D})$.
\end{definition}
Clearly a $1$-bouquet of Sasaki cones is just a Sasaki cone. Generally, the Sasaki cones in ${\mathfrak B}({\mathcal D})$ can have varying dimension.
The example of interest to us here has in its foundations in the work of Karshon \cite{Kar03} and Lerman \cite{Ler03b}. The general formulation is given in \cite{Boy10a}. We begin with a simply connected symplectic orbifold $({\mathcal B},\omega)$ such that $\omega$ defines an integral class in the orbifold cohomology group $H^2_{orb}({\mathcal B},{\mathbb Z})$. Suppose further that the class $[\omega]$ is primitive and that there are compatible almost complex structures $\hat{J}_l$ on ${\mathcal B}$ such that $(\omega,\hat{J}_l)$ is K\"ahler for each $l$ in some index set ${\mathcal A}.$ We can associate to each such $\hat{J}_l$ a conjugacy class of maximal tori in the symplectomorphism group ${\mathfrak S}{\mathfrak y}{\mathfrak m}({\mathcal B},\omega)$. Assume that this map is injective. Now form the principal $S^1$-orbibundle over ${\mathcal B}$ associated to $[\omega]$ and assume that the total space $M$ is smooth. By the orbifold version of the Boothby-Wang construction we get a contact manifold $(M,\eta)$ satisfying $\pi^*\omega=d\eta$ where $\pi:M\ra{1.5} {\mathcal B}$ is the natural orbifold projection map. We can lift a compatible almost complex structure $\hat{J}_l$ on ${\mathcal B}$ to a ${\mathcal D}$-compatible almost complex structure $J_l$ on ${\mathcal D}$. Choosing tori $T_l$ associated to $\hat{J}_l$ we can also lift these to maximal tori $\Xi\times \pi^{-1}(T_l)$ in the contactomorphism group ${\mathfrak C}{\mathfrak o}{\mathfrak n}(M,{\mathcal D})$ where ${\mathcal D}=\ker\eta$ and $\Xi$ is the circle group generated by the Reeb vector field $\xi$ of $\eta$. Suppose that there are exactly $N$ such maximal tori in ${\mathfrak C}{\mathfrak o}{\mathfrak n}(M,{\mathcal D})$, then we get an N-bouquet ${\mathfrak B}_N({\mathcal D})$ of Sasaki cones on $(M,{\mathcal D})$ which intersect in the ray of Reeb vector fields, $\xi_a=a^{-1}\xi$. If we projectivize the Sasaki cones by transverse homotheties we obtain the usual notion of bouquet, namely, a wedge product $\bigvee{\mathbb P}(\kappa({\mathcal D},\hat{J}_l))$ with base point $\xi$.
Suppose that ${\oldmathcal S}_l=(\xi,\eta,\Phi_l,g_l)$ is an extremal element of $\kappa({\mathcal D},J_l)$ for each $l\in {\mathcal A}$, that is for each $l\in {\mathcal A}$ there is an extremal representative ${\oldmathcal S}_l(t)\in {\mathcal S}(\xi,\bar{J}_l)$. Then by the openness theorem there is a nonempty open extremal set ${\mathfrak e}({\mathcal D},J_l)\subset \kappa({\mathcal D},J_l)$ containing ${\oldmathcal S}=(\xi,\eta,\Phi,g)$ for each $l\in {\mathcal A}$. This leads to
\begin{definition}\label{extbouq}
We define an {\bf extremal bouquet} ${\mathfrak E}{\mathfrak B}({\mathcal D})$ associated with the contact structure ${\mathcal D}$ to be the union $\cup_{l\in {\mathcal A}}{\mathfrak e}({\mathcal D},J_l)$ if for each $l\in {\mathcal A}$ the extremal set ${\mathfrak e}({\mathcal D},J_l)$ is non-empty. Moreover, it is called an {\bf extremal $N$-bouquet} ${\mathfrak E}{\mathfrak B}_N({\mathcal D})$ if ${\mathcal A}$ has cardinality $N$.
\end{definition}
One easily sees from the openness theorem of \cite{BGS06} that ${\mathfrak E}{\mathfrak B}({\mathcal D})$ is open in ${\mathfrak B}({\mathcal D})$. An important open problem here is to obtain a good measure of the size of the extremal sets ${\mathfrak e}({\mathcal D},J_l)$, and hence, a measure of the size of the extremal bouquet. The actual size of extremal Sasakian sets is known in very few cases, namely the standard sphere, and the Heisenberg group \cite{BGS06,Boy09}.
\section{The Main Theorems}
Let us now construct contact structures of Sasaki type on $S^2\times S^3$ and $X_\infty$. First consider the toric symplectic manifold $S^2\times S^2$ with symplectic form $\omega_{k_1,k_2}=k_1\omega_1+k_2\omega_2$ on $S^2\times S^2$ where $(k_1,k_2)$ are relatively prime integers satisfying $k_1\geq k_2$, and $\omega_1(\omega_2)$ is the standard symplectic forms on the first (second) factor of $S^2$, respectively. Let $\pi:M\ra{1.5} S^2\times S^2$ be the circle bundle corresponding to the cohomology class $[\omega_{k_1,k_2}]\in H^2(M,{\mathbb Z})$. As stated in the introduction $M$ is diffeomorphic to $S^2\times S^3$ for each such pair $(k_1,k_2)$. By the Boothby-Wang construction one obtains a contact structure ${\mathcal D}_{k_1,k_2}$ on $M$ by choosing a connection 1-form $\eta_{k_1,k_2}$ such that $\pi^*\omega_{k_1,k_2}=d\eta_{k_1,k_2}$. Justin Pati \cite{Pat09} has recently shown using contact homology that there are infinitely many such contact structures with the same first Chern class that are inequivalent.
\begin{theorem}\label{thmA}
Let $M=S^2\times S^3$ with the contact structure ${\mathcal D}_{k_1,k_2}$ described above. Let $N=\lceil \frac{k_1}{k_2} \rceil$ denote the smallest integer greater than or equal to $\frac{k_1}{k_2}$. Then for each pair $(k_1,k_2)$ of relatively prime integers satisfying $k_1\geq k_2$, there is an extremal $N$-bouquet ${\mathfrak E}{\mathfrak B}_N({\mathcal D}_{k_1,k_2})$ of toric Sasakian structures associated to the contact structure ${\mathcal D}_{k_1,k_2}$ on $S^2\times S^3$, and all the cones of the bouquet have dimension three. Furthermore, the extremal Sasakian structure corresponding to the contact 1-form $\eta_{k_1,k_2}$ has constant scalar curvature if and only if the transverse complex structure is that induced by the product complex structure on the base $S^2\times S^2$.
\end{theorem}
\begin{proof}
By Proposition \ref{SasKah} a Sasakian structure ${\oldmathcal S}=(\xi,\eta,\Phi,g)$ in $\kappa({\mathcal D}_{k_1,k_2},J)$ is extremal if and only if the induced K\"ahler structure $(\omega_{k_1,k_2},\hat{J},h)$ on $S^2\times S^2$ is extremal. Now Karshon \cite{Kar03} proved that there are precisely $\lceil \frac{k_1}{k_2} \rceil$ even Hirzebruch surfaces $S_{2m}$ that are K\"ahler with respect to the symplectic form $\omega_{k_1,k_2}$. More precisely for each $m=0,\cdots, \lceil \frac{k_1}{k_2} \rceil-1$ there is a diffeomorphism that takes the K\"ahler form $S_{2m}$ to $\omega_{k_1,k_2}$. Note that $m=0$ corresponds the product complex structure ${\mathbb C}{\mathbb P}^1\times {\mathbb C}{\mathbb P}^1$. Moreover, Calabi \cite{Cal82} has shown that for every K\"ahler class $[\omega]$ of any Hirzebruch surface $S_{n}$ there is an extremal K\"ahler metric which is of constant scalar curvature if and only if $n=0$. Furthermore, as shown by Lerman \cite{Ler03b} one can lift maximal tori of the symplectomorphism group ${\mathfrak S}{\mathfrak y}{\mathfrak m}(S^2\times S^2,\omega_{k_1,k_2})$ to maximal tori in the contactomorphism group ${\mathfrak C}{\mathfrak o}{\mathfrak n}(S^2\times S^3,{\mathcal D}_{k_1,k_2})$ (see also Theorem 6.4 of \cite{Boy10a} for the more general situation). As explained above this gives a Sasaki cone associated to each of the transverse complex structures induced by the Hirzebruch surfaces, and hence an $N$-bouquet of Sasaki cones where $N= \lceil \frac{k_1}{k_2} \rceil$. Furthermore, since each Sasaki cone has an extremal Sasakian structure, namely the one with 1-form $\eta_{k_1,k_2}$, we can apply Theorem 7.6 of \cite{BGS06} to obtain an open set of extremal Sasakian structures in each of the Sasaki cones. This gives an extremal $N$-bouquet ${\mathfrak E}{\mathfrak B}_N({\mathcal D}_{k_1,k_2})$ as defined above.
\end{proof}
There is another family of contact structures on $S^2\times S^3$ that arise as circle bundles over the non-trivial $S^2$-bundle over $S^2$ which is diffeomorphic to ${\mathbb C}{\mathbb P}^2$ blown up at a point. Following \cite{Kar03} I denote this manifold by $\widetilde{{\mathbb C}{\mathbb P}}^2$. The construction of this family also gives contact structures on the non-trivial bundle $X_\infty$. Let $E$ be the exceptional divisor and let $L$ be a projective line in $\widetilde{{\mathbb C}{\mathbb P}}^2$ that does not intersect $E$. We consider the symplectic form $\tilde{\gro}_{l,e}$ such that the symplectic areas of $L$ and $E$ are $2\pi l$ and $2\pi e$, respectively, where $(l,e)$ are relatively prime integers satisfying $l>e\geq 1$. Karshon \cite{Kar03} proved that there are precisely $\lceil \frac{e}{l-e} \rceil$ odd Hirzebruch surfaces $S_{2m+1}$ with $m=0,\cdots, \lceil \frac{e}{l-e} \rceil-1$ that are K\"ahler with respect to the symplectic form $\tilde{\gro}_{l,e}$, and again there are diffeomorphisms $\psi_m$ that take the K\"ahler forms of $S_{2m+1}$ to the form $\tilde{\gro}_{l,e}$. As occured in the proof of the previous theorem the different Hirzebruch surfaces correspond to non-conjugate maximal tori in the symplectomorphism group ${\mathfrak S}{\mathfrak y}{\mathfrak m}(\widetilde{{\mathbb C}{\mathbb P}}^2,\tilde{\gro}_{l,e})$. By Proposition 6.3 and Theorem 6.4 of \cite{Boy10a} these can be lifted to non-conjugate tori in the contactmorphism group ${\mathfrak C}{\mathfrak o}{\mathfrak n}(M,\tilde{\cald}_{l,e})$ where $M$ is the circle bundle over $\widetilde{{\mathbb C}{\mathbb P}}^2$ corresponding to the cohomology class $[\omega_{l,e}]$ and $\tilde{\cald}_{l,e}$ is the contact structure induced by a connection 1-form $\eta_{l,e}$ satisfying $\pi^*\omega_{l,e}=d\eta_{l,e}$. I assume that $l$ and $e$ are relatively prime positive integers. Now in order to identify the diffeomorphism type of $M$ it is enough by the Barden-Smale classification of simply connected 5-manifolds to compute the mod 2 reduction of the first Chern class $c_1(\tilde{\cald}_{l,e})$. For this we pullback $c_1(\widetilde{{\mathbb C}{\mathbb P}}^2)$ to $M$ and use the relation $[\pi^*\tilde{\gro}_{l,e}]=0$. Now $c_1(\widetilde{{\mathbb C}{\mathbb P}}^2)=2\alpha_E+\alpha_L$ where $\alpha_E(\alpha_L)$ is Poincar\'e dual of the divisor class $E(L)$, respectively \cite{GrHa78}. We find $c_1(\tilde{\cald}_{l,e})=\bigl(l-2e\bigr)\gamma$ where $\gamma$ is a generator of $H^2(M,{\mathbb Z})\approx {\mathbb Z}$. It follows that $M$ is $S^2\times S^3$ if $l$ is even and $X_\infty$ if $l$ is odd. Then, using Calabi \cite{Cal82}, just as in the proof of Theorem \ref{thmA} we arrive at
\begin{theorem}\label{thmB}
Let $M$ be the circle bundle over $\widetilde{{\mathbb C}{\mathbb P}}^2$ with the contact structure $\tilde{\cald}_{l,e}$ as described above, and let $N=\lceil \frac{e}{l-e} \rceil$. Then for each pair $(l,e)$ of relatively prime integers satisfying $l>e\geq 1$, there is an extremal $N$-bouquet ${\mathfrak E}{\mathfrak B}_N(\tilde{\cald}_{l,e})$ of toric Sasakian structures associated to the contact structure $\tilde{\cald}_{l,e}$, and the cones of the bouquet all have dimension three. Furthermore, if $l$ is even $M=S^2\times S^3$, whereas, if $l$ is odd $M=X_\infty$ and in either case the extremal Sasakian structure corresponding to the contact 1-form $\eta_{l,e}$ does not have constant scalar curvature.
\end{theorem}
\begin{remark}
If the pairs of integers $(k_1,k_2)$ and $(l,e)$ are not relatively prime, but $(k_1,k_2)=(l,e)=n$, similar results to Theorems \ref{thmA} and \ref{thmB} hold for the quotient manifolds $(S^2\times S^3)/{\mathbb Z}_n$ and $X_\infty/{\mathbb Z}_n$.
\end{remark}
\section{Concluding Remarks}
Recall \cite{BG05} that a Sasakian structure ${\oldmathcal S}=(\xi,\eta,\Phi,g)$ is said to be {\it positive (negative)} if its basic first Chern class $c_1({\mathcal F}_\xi)$ can be represented by a positive (negative) definite $(1,1)$-form. It is {\it null} if $c_1({\mathcal F}_\xi)=0$, and {\it indefinite} otherwise. The following result follows directly from Theorem 8.1.14 of \cite{BG05} and Proposition 4.4 of \cite{BGS06}.
\begin{lemma}\label{CRauttype}
Let ${\oldmathcal S}=(\xi,\eta,\Phi,g)$ be a Sasakian structure with underlying CR structure $({\mathcal D},J)$. Suppose that $\dim \kappa({\mathcal D},J)>1$. Then the type of ${\oldmathcal S}$ is either positive or indefinite.
\end{lemma}
In a forthcoming work \cite{BoPa10} it will be seen that generally the type is not an invariant of the Sasaki cone. Here it is easy to see from the constructions above that the type is not an invariant of the bouquet. For example consider the contact structure ${\mathcal D}_{5,1}$ on $S^2\times S^3$. There are five Sasaki cones $\kappa({\mathcal D}_{5,1},J_{2m})$ where $m=0,\cdots,4$. We can label the corresponding extremal Sasakian structures as ${\oldmathcal S}_{2m}=(\xi,\eta_{5,1},\Phi_{2m},g_{2m})$ where $J_{2m}=\Phi_{2m} |_{{\mathcal D}_{5,1}}$ and $g_{2m}= d\eta_{5,1}\circ(\Phi_{2m}\otimes {\mathchoice {\rm 1\mskip-4mu l} {\rm 1\mskip-4mu l)+\eta_{5,1}\otimes \eta_{5,1}$. The Sasakian structure ${\oldmathcal S}_0$ is positive and has constant scalar curvature; whereas, the others ${\oldmathcal S}_{2m}$ with $m=1,2,3,4$ are indefinite with non-constant scalar curvature. Similarly, the contact structure $\tilde{\cald}_{5,4}$ on $X_\infty$ has four Sasaki cones $\kappa(\tilde{\cald}_{4,1},J_{2m+1})$ with $m=0,1,2,3$. The corresponding extremal Sasakian structures are ${\oldmathcal S}_{2m+1}=(\xi,\tilde{\eta}_{4,1},\Phi_{2m+1},g_{2m+1})$ where $J_{2m+1}=\Phi_{2m+1} |_{\tilde{\cald}_{4,1}}$ and $g_{2m+1}= d\tilde{\eta}_{5,1}\circ(\Phi_{2m+1}\otimes {\mathchoice {\rm 1\mskip-4mu l} {\rm 1\mskip-4mu l)+\tilde{\eta}_{4,1}\otimes \tilde{\eta}_{4,1}$. Again ${\oldmathcal S}_1$ is positive, and the others are indefinite, but now they are all extremal of non-constant scalar curvature.
The case $(l,e)=(2,1)$ and $m=0$ presents an interesting case even though the Sasaki bouquet degenerates to a single Sasaki cone. Here we have $c_1((\tilde{\cald}_{2,1},J_0))=0$ and $M=S^2\times S^3$. So by Calabi the induced Sasaki structure over the complex manifold $\widetilde{{\mathbb C}{\mathbb P}}^2$ has an extremal Sasaki metric with non-constant scalar curvature. However, it was shown in \cite{MaSp06} that
the contact structure $\tilde{\cald}_{2,1}$ admits an irregular Sasaki-Einstein structure, which, of course, is extremal with constant scalar curvature. In this case it would be interesting to see how big the extremal set ${\mathfrak e}(\tilde{\cald}_{l,e},J_0)$ is. This is actually a special case of a much more general result of Futaki, Ono, and Wang \cite{FOW06} which says that for any positive toric Sasakian structure with $c_1({\mathcal D})=0$ there is a deformation to a Reeb vector field in the Sasaki cone whose Sasakian metric is Sasaki-Einstein. This discussion begs the questions:
\noindent {\it Does every toric contact structure of Reeb type on $S^2\times S^3$ or $X_\infty$ have a constant scalar curvature Sasakian metric somewhere in its Sasaki cone?, or more generally for any toric contact structure of Reeb type?}
Another observation comes from the well known fact \cite{MoKo06} that for each $m>0$ there is a complex analytic family of even Hirzebruch surfaces $S_{2l}(t)$ satisfying $S_{2l}(t)\approx S_{2l}$ for $t\neq 0$ and $m>l$, but that $S_{2l}(0)\approx S_{2m}.$ This implies that the moduli space of complex structures on $S^2\times S^2$ is non-Hausdorff. A similar result holds for the moduli space of complex structures on $\widetilde{{\mathbb C}{\mathbb P}}^2$. These results imply that the moduli space of extremal Sasakian structures with underlying contact structures ${\mathcal D}_{k_1,k_2}$ on $S^2\times S^3$ or $\tilde{\cald}_{l,e}$ on $X_\infty$ is non-Hausdorff as well.
Note added: The first question above has been recently answered in the affirmative by \'Eveline Legendre \cite{Leg10}. Moreover, she shows that if $k_1>5k_2$ the toric contact structure ${\mathcal D}_{k_1,k_2}$ on $S^2\times S^3$ has two distinct rays in the Sasaki cone with constant scalar curvature Sasakian metrics. So unlike the Sasaki-Einstein case \cite{CFO07,FOW06} constant scalar curvature rays are not unique in a Sasaki cone.
\section*{Acknowledgements}
I would like to thank V. Apostolov for many discussions about extremal K\"ahler metrics.
\def$'$} \def\cprime{$'$} \def\cprime{$'$} \def\cprime{$'${$'$} \def$'$} \def\cprime{$'$} \def\cprime{$'$} \def\cprime{$'${$'$} \def$'$} \def\cprime{$'$} \def\cprime{$'$} \def\cprime{$'${$'$} \def$'$} \def\cprime{$'$} \def\cprime{$'$} \def\cprime{$'${$'$}
\def$'$} \def\cprime{$'$} \def\cprime{$'$} \def\cprime{$'${$'$} \def$'$} \def\cprime{$'$} \def\cprime{$'$} \def\cprime{$'${$'$} \def$'$} \def\cprime{$'$} \def\cprime{$'$} \def\cprime{$'${$'$} \def$'$} \def\cprime{$'$} \def\cprime{$'$} \def\cprime{$'${$'$}
\def$''$} \def\cprime{$'$} \def\cprime{$'${$''$} \def$'$} \def\cprime{$'$} \def\cprime{$'$} \def\cprime{$'${$'$} \def$'$} \def\cprime{$'$} \def\cprime{$'$} \def\cprime{$'${$'$}
\providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
\providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR }
\providecommand{\MRhref}[2]{%
\href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2}
}
\providecommand{\href}[2]{#2}
|
\section{The \textit{Chandra}\ View of Young SNRs}
\subsection{The Power of the Data}
The excellent quality of the data generated by \textit{Chandra}\ for bright SNRs in our Galaxy is showcased by the false color images
that have become perhaps the mission's most recognizable and widespread visual result (Figure 1). For a SNR
with an angular diameter of $6 ^\prime$ like Cas A, \textit{Chandra}\ can resolve more than $10^{5}$ individual regions. Deep
exposures of bright objects usually provide enough photon statistics to obtain a good spectrum from most of these
regions at the moderate spectral resolution of the ACIS CCD detectors ($E / \Delta E \approx 10-60$). This usually
allows to detect K-shell emission from abundant elements like O, Si, S, Ar, Ca, and Fe.
\begin{figure*}[t]
\begin{center}
\includegraphics[width=\textwidth]{tycho_kepler_casa.eps}
\caption{Three-color images generated from deep \textit{Chandra}\ exposures of the Tycho (left), Kepler (center), and Cas A
(right) SNRs. The details vary for each image, but red usually corresponds to low energy X-rays around the Fe-L
complex ($\sim$ 1 keV and below), green to mid-energy X-rays around the Si K blend ($\sim$ 2 keV), and blue to
high energy X-rays in the 4-6 keV continuum bewteen the Ca K and Fe K blends. Images are not to scale: Tycho is
$\sim8\, ^\prime$ in diameter, Kepler is $\sim4\, ^\prime$, and Cas A is $\sim6\, ^\prime$. Total exposure times are
$150$, $750$, and $1000$ ks. Images courtesy of the \textit{Chandra}\ X-ray Center; data originally published in
\cite{warren05:Tycho}, \cite{reynolds07:kepler}, and \cite{hwang04:CasA_VLP}.}\label{fig-1}
\end{center}
\end{figure*}
\subsection{Basic Concepts of NEI Plasmas}
The density of the plasma inside SNRs is low enough for the ages of young objects like Cas A or Tycho to be smaller than
the ionization equilibrium timescale. The X-ray emitting plasma heated by the shocks is therefore in a state of
nonequilibrium ionization, or NEI \cite{itoh77:SNRs_NEI}. This means that the ionization state of any given fluid
element is determined not only by its electron temperature $T_{e}$, but also by its ionization timescale $n_{e}t$, where
$n_{e}$ is the electron number density and $t$ is the time since shock passage. The thermal X-ray spectrum from a young
SNR is thus intimately related to its dynamic evolution through the individual densities and shock passage times of each
fluid element. This has important implications for the ejecta emission, because of the large differences in chemical
composition across the SN debris, and the fact that the electron pool is completely dominated by the contributions from
high-Z elements, making $n_{e}$ a strong function of the ionization state \cite{hamilton84:ejecta}.
Under these circumstances, the quantitative analysis of X-ray spectra from young SNRs becomes a challenging endeavor. In
order to derive magnitudes that are relevant to SN physics, like the kinetic energy $E_{k}$ or the ejected mass of each
chemical element, it is necessary to build a hydrodynamic model of the entire SNR. One must know when each fluid element
was shocked, what its chemical composition is, how much of the SN ejecta is still unshocked at the present time,
etc. Individual fitting of each magnitude becomes impractical, because they are all related to each other, and in order
to understand the X-ray spectrum of a particular SNR, one has to understand of the object as a whole. Further
complications stem from two sources. One is a technical, but important, problem: the uneven quality of atomic data in
X-ray emission codes for NEI plasmas \cite{borkowski01:sedov,badenes06:tycho}. The other is of a more fundamental
nature: the large uncertainties in the physics of collisionless shocks, in particular the amount of ion-electron
temperature equilibration at the shock transition \cite{ghavamian07:shock_equilibration,heng09:balmer_shocks} and the
impact of cosmic ray acceleration on the dynamics of the plasma \cite{decourchelle00:cr-thermalxray}.
Most of the analysis of the X-ray emission from SNRs is done following one of two radically different approaches. The
first approach is to forgo the complex interaction between SNR dynamics and X-ray spectrum, and fit the emission of the
entire SNR or individual regions using ready-made spectral models. The most widely used are plane-parallel shock NEI
models \cite{hughes00:E0102,borkowski01:sedov}, which fit the spectra by varying $T_{e}$, some parameter related to
$n_{e}t$, and a set of chemical abundances. This has the advantages of simplicity and flexibility, but the number of
free parameters is large, the quality of the fits is often poor, and the results are hard to interpret in the framework
of SN physics and progenitor scenarios. The other approach is to model the full HD evolution of the SNR \textit{ab
initio}, starting from a grid of SN explosion models and ISM or CSM configurations, calculate the NEI processes in the
shocked plasma, and produce a set of synthetic X-ray spectra that can then be compared to the observations. This
approach is less flexible, and usually does not allow for spectral fits in the usual sense (i.e., based on the $\chi^2$
statistic), but it is considerably more powerful in that the observations and the physical scenarios that are being
tested can be compared in a more direct way. The first HD+NEI models for SNRs were built to analyze the data from early
X-ray missions like \textit{Einstein} and \textit{EXOSAT} (e.g,
\cite{hughes85:nei,hamilton86:SN1006,borkowski94:kepler}). Modern efforts produced for \textit{Chandra}\ and \textit{XMM-Newton}\ data can be found
in \cite{badenes03:xray,badenes05:xray} and \cite{sorokina04:typeIasnrs}.
\section{Type Ia SNe and Their SNRs}
\subsection{Open Issues in Type Ia SNe}
Type Ia SNe are believed to be the thermonuclear explosion of a C+O WD that is destabilized when its mass becomes close
to the Chandrasekhar limit by accretion of material from a binary companion. After the central regions ignite, the
burning front propagates outwards, consuming the entire star and leading to a characteristic ejecta structure, with
$\sim 0.7\,\mathrm{M_{\odot}}$ of Fe-peak nuclei (mostly the $^{56}$Ni that powers the light curve) in the inside and
about an equal amount of intermediate mass elements (mostly Si, S, Ar, and Ca) in the outside. The amount of $^{56}$Ni
can be as high as $\sim 1\,\mathrm{M_{\odot}}$ in the brightest Type Ia SNe, and as low as $\sim
0.3\,\mathrm{M_{\odot}}$ in the dimmest Type Ia SNe, but the large scale stratification seems to be retained in all cases
\cite{mazzali07:zorro}. State-of-the-art three-dimensional explosion models still cannot reproduce these basic features
in a self-consistent way, mostly because hydrodynamic instabilities tend to destroy the stratification of nucleosynthetic
products on very short timescales \cite{kasen09:SNIa_diversity}. This relatively simple structure of Type Ia SN ejecta
is responsible for the uniformity of the light curves and spectra that makes them useful as distance indicators for
cosmology. The still poorly understood details of the multi-dimensional explosion mechanism might provide the diversity
within the uniformity, leading to bright and dim events \cite{kasen09:SNIa_diversity}. The fact that bright Ia SNe
appear to explode preferentially in star-forming galaxies \cite{gallagher05:chemistry_SFR_SNIa_hosts} suggests some
connection between the properties of the progenitors and the explosion physics, but since the identity of the
progenitors has never been clearly established, this connection remains mysterious
\cite{maoz08:fraction_intermediate_stars_Ia_progenitors}. Depending on the nature of the WD companion, SN Ia progenitors
are divided into single degenerate (the companion is a normal star, \cite{hachisu96:progenitors}) and double degenerate
(the companion is another WD \cite{iben84:typeIsn,webbink84:DDWD_Ia_progenitors}). Finding clear evidence supporting one
of these two scenarios has become a central problem in stellar astrophysics, because unknown evolutionary effects
associated with the progenitors may introduce systematic trends that could limit the precision of cosmological
measurements based on SN Ia \cite{howell09:SN_white_paper}.
\subsection{Type Ia SNRs}
Most (but not all) Type Ia SNRs reflect the relative uniformity and simplicity of their birth events. Even a cursory
glance at the \textit{Chandra}\ images of Tycho (Figure 1, \cite{warren05:Tycho}) or SN1006 \cite{cassam-chenai08:SN1006} reveals
strikingly symmetrical objects, without large anisotropies in the ejecta emission, in marked contrast to the turbulent
nature of CC SNRs. The morphology and X-ray spectra of SNRs with known ages and a well-established Type Ia
classification also indicate that the progenitors do not modify their surroundings in a strong way - in particular,
there is no evidence for large wind-blown cavities around the explosion sites \cite{badenes07:outflows}. This is
relevant because the current paradigm for Type Ia progenitors in the single degenerate channel predicts fast, optically
thick outflows from the WD surface \cite{hachisu96:progenitors} that would leave behind such cavities. We have to
conlcude that either most Type Ia SNe do not have single degenerate progenitors, or these fast outflows are not always
present. In fact, the dynamical and spectral properties of young Type Ia SNRs like Tycho or 0509$-$67.5\ are consistent with an
interaction with the warm phase of the ISM \cite{badenes06:tycho,badenes08:0509}, and so is the detailed morphology of
the blast wave in objects where it can be studied with sufficient resolution, like SN1006 \cite{raymond07:SN1006_shocks}
A notable exception to this is the Kepler SNR. Although the recent deep (750 ks) \textit{Chandra}\ exposure of Kepler left little
doubt that the SNR is of Type Ia \cite{reynolds07:kepler}, its morphology shows a strong bilateral asymmetry (Figure
1), and optical observations reveal clear signs of an interaction with a dense, N-enriched CSM
\cite{blair91:kepler_optical}. This suggests that either the progenitor of Kepler's SN or its binary companion might
have been relatively massive, creating a bow-shock shaped CSM structure as the system lost mass and moved against the
surrounding ISM \cite{bandiera87:kepler}. This complex CSM interaction makes HD+NEI modeling of the Kepler SNR
challenging, and the estimation of the fundamental SN parameters difficult. Most of the detailed models for Kepler were
published when the SNR was widely believed to be of CC origin
\cite{bandiera87:kepler,borkowski94:kepler}. It is important that these models be revised in light of
the secure Ia origin of the SNR to determine the mass of $^{56}$Ni synthesized and the pre-explosion mass-loss rate from
the progenitor.
\begin{figure}
\begin{center}
\includegraphics[scale=0.87,angle=90]{Tycho_Shocks.eps}
\caption{HD+NEI models for the Tycho SNR: Inferred distance as a function of the ambient medium
density $\rho_{AM}$ obtained by matching the angular sizes of the reverse shock (RS, dashed plots) and contact
discontinuity (CD, solid plots) to a bright Ia SN model (DDTa, blue plots), and a dim Ia SN model (DDTg, red
plots). The values of $\rho_{AM}$ in the HD+NEI models of \cite{badenes06:tycho} that provide the best match to
the X-ray emission from the SN ejecta are indicated by the striped vertical band. The value of $\rho_{AM}$
required by the CR-modified shock models of \cite{cassam-chenai07:tycho} is indicated by a vertical dash-dotted
line. The estimated value of the distance $D$ from \cite{smith91:six_balmer_snrs} is indicated by the horizontal
solid line, with the dotted horizontal lines marking the upper and lower limits.}\label{fig-2}
\end{center}
\end{figure}
The HD+NEI modeling approach has had more success in objects with simpler dynamics, like the Tycho SNR. One-dimensional
HD+NEI models with a uniform ambient medium can reproduce the fundamental features of both the dynamics and the
integrated X-ray spectrum of Tycho, provided that the ejecta is stratified, with an Fe content similar to what would be
expected from a SN Ia of normal brightness, and the ambient medium density is $\rho_{AM} \approx 2 \times
10^{-24}\,\mathrm{g\,cm^{-3}}$ (\cite{badenes06:tycho}, Figures 2 and 3). This density estimate
merits a few comments. Although it is clear that cosmic ray acceleration has a strong impact on the dynamics of the
forward shock \cite{decourchelle00:cr-thermalxray,ellison04:hd+cr}, the fact remains that the microphysics of the
process is not well understood, and cosmic ray backreaction is usually not included in full HD+NEI models for the
ejecta emission in SNRs (see \cite{patnaude09:DSA_NEI}). Instead, the assumption is made that the dynamics of the
shocked ejecta are not severely affected by cosmic ray acceleration at the blast wave, which seems reasonable in light
of the properties of the reverse shock (see discussion in Section 3 of \cite{badenes06:tycho}). In any case, it is
important to emphasize that the value of $\rho_{AM}$ required by the ejecta emission in the Tycho SNR also reproduces
the angular sizes of the reverse shock and contact discontinuity at the correct distance to the SNR (Figure 2). This
basic validation of the underlying dynamic model is essential to the HD+NEI approach, and strengthens the confidence on
the conclusions drawn about the SN explosion.
\begin{figure}
\begin{center}
\includegraphics[scale=0.65]{DDTc_Tycho.eps}
\caption{HD+NEI models for the Tycho SNR: Best-fit HD+NEI model to the total X-ray spectrum. The SN model is DDTc, a
normal SN Ia model that synthesizes $0.74\,\mathrm{M_{\odot}}$ of $^{56}$Ni, interacting with an AM of $\rho_{AM}
= 2 \times 10^{-24}\,\mathrm{g\,cm^{-3}}$. The contribution of the nonthermal emission from the blast wave is
indicated by the dotted plot, the rest is thermal emission from the shocked ejecta (figure from
\cite{badenes06:tycho}).}\label{fig-3}
\end{center}
\end{figure}
The stratification of the ejecta in the Tycho SNR is required by the fact that the $n_{e}t$ of Fe in the X-ray spectrum
is lower than that of Si, which can be explained naturally if most of the Si is shocked earlier and at a higher density
than most of the Fe (see also \cite{kosenko06:Tycho}). This argument is completely independent from the constraints on
the ejecta structure obtained with optical spectroscopy of Type Ia SNe, and constitutes a clear sign that a supersonic
burning front (i.e., a detonation) must have been involved in the physics of the explosion at some stage. The most
succesful SN model found by \cite{badenes06:tycho} is a one-dimensional delayed detonation with $E_{k}=1.2
\times10^{51}$ erg that synthesizes $0.74\,\mathrm{M_{\odot}}$ of $^{56}$Ni (Figure 3) - in other words, a Type Ia SN of
normal brightness, consistent with the historical records \cite{ruiz-lapuente04:TychoSN}.
\begin{figure}
\begin{center}
\includegraphics[scale=1.3]{light_echo_0509.eps}
\caption{Direct comparison between SN and SNR spectroscopy for the LMC SNR 0509$-$67.5: The LE is well matched only by spectra
of bright Type Ia SNe like SN 1998es, SN 1999aa, and SN 1999dq (figure from \cite{rest08:0509}). }\label{fig-4}
\end{center}
\end{figure}
One largely unexpected, but very important, development in this field has been the possibility to calibrate the results
obtained from the analysis of the ejecta emission in SNRs using the light echoes from ancient SNe. These light echoes
were originally discovered in the LMC \cite{rest05:LMC_light_echoes}, and they are important for two reasons. First,
they can provide a reliable and independent age estimate for objects that previously had none, which is a crucial
ingredient to build HD models (see e.g. \cite{badenes07:outflows}). Second, and most important, the spectroscopy of the
light echoes can be used to determine the type (CC vs. Ia) and subtype (e.g., bright Ia or dim Ia) of ancient SNe. This
technique was demonstrated by \cite{rest08:0509}, who showed that SNR 0509$-$67.5\ in the LMC was originated by a bright,
$^{56}$Ni-rich Type Ia SN. At the same time, \cite{badenes08:0509} revisited the archival observations of this SNR and
found that an HD+NEI model based on a bright SN Ia explosion reproduces both the X-ray emission from the shocked ejecta
and the dynamics of the SNR (see Figures 4 and 5). This agreement between two completely independent
techniques applied to the same object has now been extended to historical SNe in our own Galaxy, with the discovery of
light echoes from Tycho and Cas A \cite{rest08:CasA_Tycho_LEs}. Spectroscopic analysis of the light echo from Tycho
\cite{krause08:tycho} confirmed that SN1572 was indeed a Type Ia SN of normal brightness, a spectacular validation of
the previous result obtained with HD+NEI models of the SNR \cite{badenes06:tycho}.
\begin{figure}
\begin{center}
\includegraphics[scale=0.87,angle=90]{0509_3models.eps}
\caption{Direct comparison between SN and SNR spectroscopy for the LMC SNR 0509$-$67.5: The X-ray spectrum of the SNR can
only be reproduced with a bright Type Ia SN models that synthesizes $0.97\,\mathrm{M_{\odot}}$ of $^{56}$Ni
(figure adapted from \cite{badenes08:0509}).}\label{fig-5}
\end{center}
\end{figure}
The emergence of the X-ray observations of SNRs as well-established probes for Type Ia SNe has important implications
for SN research. Because SNRs are nearby objects, the context of the exploded star can be studied in great detail, and
relationships between specific explosion properties and progenitor scenarios can be tested \textit{in situ}. One
powerful constraint on the single degenerate scenario for SN Ia progenitors, for example, is the presence of the donor
star, which should survive the explosion \cite{canal01:companions}. So far, the search for this star in the Tycho SNR
has produced inconclusive results \cite{ruiz-lapuente04:Tycho_Binary,hernandez09:Tycho_G,kerzendorf09:Tycho_G}, but
further work in this and other SNRs should eventually clarify the identity of the progenitor of at least one SN Ia
explosion with known properties. Another recently opened line of research is the possibility to measure the metallicity
of SN Ia progenitors \textit{directly}, using Mn and Cr lines in the X-ray spectrum of their SNRs. This technique yields
a supersolar metallicity for the progenitor of Tycho \cite{badenes08:mntocr}, a measurement that will soon be extended
to other objects like Kepler. Even a small census of metallicities obtained in this way will allow us to validate the
correlations between the metallicity of the host galaxy and the SN Ia brightness found in SN surveys
\cite{cooper09:metallicity_bias_SNIa}. Finally, the SNR population in the Magellanic Clouds has great potential to
constrain Type Ia SN progenitor models, because the resolved stellar population in these nearby galaxies can be studied
in great detail. SNR 0509$-$67.5, for instance, which was originated by a bright Type Ia SN, is embedded in a relatively old,
metal-poor stellar population, while SNR N103B, which has signs of a CSM interaction similar to Kepler
\cite{lewis03:N103B}, is associated with a much younger, metal-rich population \cite{badenes09:SNRs_LMC}. This suggests
that the progenitor of SNR N103B might have been relatively young and massive, and might have modified its surroundings
through some kind of mass-loss process.
\section{Core Collapse SNe and Their SNRs}
\subsection{Open Issues in Core Collapse SNe}
CC SNe mark the final events in the lives of massive ($\gtrsim 8\,\mathrm{M_{\odot}}$) stars. After nuclear fuel is
exhausted in the inner regions, the stellar core collapses to form a neutron star or a black hole. The layers of the
star that do not end up in the central compact object bounce and start propagating outwards, but at present it is
unclear exactly how the gravitational collapse of the core becomes an explosion \cite{janka07:CCSN_Review}. At the
writing of this review, the most sophisticated multi-D simulations of CC SNe still fail to explode. Although the
identity of the progenitors of CC SNe is well-established, with several identifications in pre-explosion images
\cite{smartt09:CCSN_progenitors}, the details of which specific kinds of massive stars lead to
which specific subtypes of CC SNe (e.g. Type Ib/c, Type IIL, etc.) are still under debate. Most of these uncertainties
stem from our imperfect knowledge of key processes in stellar evolution, like mass-loss mechanisms or the role played by
binarity.
\subsection{Core Collapse SNRs}
Just like Type Ia SNRs, the basic morphology of CC SNRs reflects the fundamental characteristics of their birth
events. The \textit{Chandra}\ images of young CC SNRs like Cas A \cite{hwang04:CasA_VLP} or SNR G292.0$+$1.8 \cite{park07:G292}
reveal a complex and turbulent structure, with marked asymmetries related both to the SN ejecta and the CSM. In contrast
to Type Ia SNRs, many CC SNRs show prominent optical emission from radiatively cooled ejecta, the hallmark of an
interaction with a relatively dense medium.
The poster child for young, ejecta-dominated CC SNRs is Cas A, an object that has played a central role in the history
of \textit{Chandra}. The first scientific result from the mission was the discovery of the central compact object
\cite{tananbaum99:CasA}, now thought to be a neutron star with a C atmosphere \cite{wynn09:CasA_NS_Catm}. The first
refereed publication based on \textit{Chandra}\ data was also devoted to Cas A \cite{hughes00:casA}. This paper discusses the
overturn of nucleosynthetic products in a large region in the SE of the SNR, where Fe-rich material can be seen clearly
ahead of Si-rich material, indicating that some of the layers deep into the onion-skin structure of the exploding star
have overtaken regions dominated by lower-Z elements (Figure 1, however, see \cite{delaney10:CasA3D} for an alternative
interpretation). Over the years, more and more has been learned about the Cas A SN and its progenitor. In 2003,
\cite{chevalier03:CasA} used the positions of the fluid discontinuities, the presence of high-velocity H, and the extent
of the clumpy photoionized pre-SN wind to infer that the Cas A explosion must have been of Type IIn or IIb SN. This
`prediction of the past', similar to the one made for the Tycho SNR, was also confirmed by the light echo, which has a
spectrum very similar to that of the Type IIb SN1993J \cite{krause08:CasA_light_echo}.
\begin{figure}[t]
\begin{center}
\includegraphics[width=.42\textwidth]{casa2_3d.eps}
\caption{The Cas A SNR: Three-dimensional visualization of the distribution of ejecta,
built by T. Delaney using Doppler shift mapping of multi-wavelength data \cite{delaney10:CasA3D}: green is
X-ray emitting Fe; yellow is X-ray, optical and infrared emitting Ar and Si; red is infrared emitting unshocked
ejecta; the pink dot represents the compact object. }\label{fig-6}
\end{center}
\end{figure}
The morphology of the ejecta emission from Cas A is very rich (\cite{willingale02:CasA}, Figures 1 and 6), but
perhaps the most striking feature in the X-ray images is the jet/counterjet structure that runs from the SW to the NE
\cite{hwang04:CasA_VLP}. In order to understand the energetics of this feature and the role that it played in the
explosion, \cite{laming03:X-ray_knots_CasA,hwang03:casa-Fe} used HD+NEI simulations to interpret the fits to small X-ray
bright knots in the SN ejecta (see Figure 7). While it is clear that there was a significant local deposition of energy
around the jet region \cite{laming03:X-ray_knots_CasA}, this was probably not enought to drive the entire explosion
\cite{laming06:CasA_polar_regions} (see also \cite{wheeler08:CasA_Shape}). It turns out that the proper motion of the
compact object is not aligned with the jet axis, as one would expect in a jet-driven explosion, but perpendicular to it,
along a mysterious gap in the high-velocity ejecta revealed by \textit{HST} images of optical knots
\cite{fesen06:CasA_HST}. The expansion of these knots is consistent with an explosion date close to 1680
\cite{thorstensen01:CasA_expansion,fesen06:CasA_HST}, but the SN was not nearly as bright as other historical events
like Tycho or Kepler. Given the reddening towards the line of sight of the SNR, this suggests a low ejected mass of
$^{56}$Ni, $\sim0.1\,\mathrm{M_{\odot}}$ \cite{eriksen09:CasA}. Cas A is also one of the few CC SNRs that is young
enough to detect the decay products of $^{44}$Ti and infer an ejected mass for this isotope
($\sim10^{-4}\,\mathrm{M_{\odot}}$, \cite{vink01:CasA-44Ti}).
\begin{figure}
\begin{center}
\includegraphics[width=.49\textwidth]{ejecta_o2.eps}
\caption{The Cas A SNR: Constraints on the power-law index of the ejecta structure obtained by comparing the fitted
$T_{e}$ and $n_{e}t$ values in small X-ray knots to HD+NEI models. The lines represent models appropriate for Cas
A, with $M_{ej}=2\,\mathrm{M_{\odot}}$, $E_{k}=2\times10^{51}$ erg and different power-law indexes. Figure from
\cite{laming03:X-ray_knots_CasA}. }\label{fig-7}
\end{center}
\end{figure}
Several authors have used this wealth of observational information to build detailed models for the progenitor of the
Cas A SN. According to \cite{young06:CasA_Progenitor}, all the observational constraints from the SNR can be matched by
a progenitor scenario where a 15-25 $\mathrm{M_{\odot}}$ star loses its H envelope to a binary interaction and undergoes
an energetic explosion in a `stripped' state. This view has been challenged by later works, who favor a slightly more
massive progenitor \cite{perez09:CasA_Progenitor}. A particularly controversial issue is whether the progenitor went
through a Wolf-Rayet phase before exploding, excavating a small cavity inside the dense $\rho \propto r^{-2}$ CSM that
is being overrun by the blast wave today \cite{schure08:CasA,vanveelen09:noWR_CasA,hwang09:CasA_CSM}.
This brief (and incomplete) summary of what we have learned about the Cas A SN from its SNR exemplifies the great value
of deep observations of nearby SNRs. Although the explosion itself was never observed with the conventional tools used
to study distant SNe, the birth event of the Cas A SNR might be the best-studied demise of a massive star after that of
SN 1987A.
\section{Conclusions and Future Perspectives}
This brief review on the X-ray observations of SNRs has been written with two main goals. The first is to illustrate the
power of SNR studies as probes of the SN phenomenon. The unique capabilities of modern X-ray satellites like \textit{Chandra}\ offer
a radically different view of the explosions and their progenitors, a vision that is independent of, and complementary
to, the conventional studies of extragalactic SNe at optical wavelengths. The X-ray data sets assembled for well-known
objects like Tycho or Cas A will be a lasting legacy of the \textit{Chandra}\ mission, and they represent the most detailed picture
of the structure of SN ejecta currently available at any wavelength. The recent discovery that the statistical
properties of the ejecta emission resolved by \textit{Chandra}\ can be used to distinguish CC from Type Ia SNRs
\cite{lopez09:typing_SNRs} is a powerful example of the potential of these data. It is therefore crucial that the
campaign of X-ray observations of SNRs with \textit{Chandra}\ and other satellites continue, and that deep exposures be completed for
all the objects that do not have them. The discovery of new young SNRs, while difficult, is definitely possible, as
illustrated by the $\sim100$ yr old Galactic SNR G1.9$+$0.3 \cite{reynolds09:G1.9+0.3}, and should be pursued in
parallel to the study of well-known ones.
The second goal of this review is to emphasize the importance of careful modeling for the analysis of the X-ray
spectra of SNRs. Because of the NEI character of the shocked plasma in SNRs, it is extremely difficult to interpret
their X-ray spectra in terms of magnitudes that can be used effectively to constrain explosion physics and progenitor
scenarios. A robust, quantitative analysis usually requires putting together a dynamical model for the SNR - in other
words, the X-ray emission of a SNR cannot be interpreted without understanding the object as a whole, at least to some degree. In
this context, one-dimensional HD+NEI models have been successful in recovering the fundamental properties of SN
explosions from the spatially integrated X-ray spectra of Type Ia SNRs, and validation of the technique through the
spectroscopy of SN light echoes has been possible in a few cases.
These results are certainly encouraging, but much remains to be done before the full potential of the X-ray observations
of SNRs is realized. In order to take advantage of the spatially resolved spectroscopy capabilities of \textit{Chandra}, it is
necessary to build multi-dimensional HD+NEI SNR models. This line of research has great promise as a benchmark for
multi-dimensional SN explosion models, which hold the key to fundamental processes like the physical mechanism
responsible for CC SNe \cite{janka07:CCSN_Review} or the origin of the diversity within Type Ia SNe
\cite{kasen09:SNIa_diversity}.
Present and future X-ray observations of SNRs present many opportunities for SN research. These opportunities are not
devoid of challenges, but the excellent quality of the data obtained with \textit{Chandra}\ and other satellites fully warrant the
effort required to meet them. Because of the unique view of the SN phenomenon that they offer, SNRs should play a
central role in shaping our understanding of SN explosions.
\begin{acknowledgments}
I want to thank Kimberly Arcand and Peter Edmonds for assistance with the \textit{Chandra}\ images. I also want to express my
gratitude to all the mebers of the SN and SNR community with whom I have shared the excitement about the results of
\textit{Chandra}\ and other X-ray satellites.
\end{acknowledgments}
|
\section{Introduction} \label{sec:intro}
The Fermat-Weber center of a measurable planar set $Q$ with positive area
is a point in the plane that minimizes the average distance to the points in $Q$.
Such a point is the ideal location for a base station (e.g., fire station
or a supply station) serving the region $Q$, assuming the region has
uniform density. Given a measurable set $Q$ with positive area and a
point $p$ in the plane, let $\mu_Q(p)$ be the average distance between
$p$ and the points in $Q$, namely,
$$ \mu_Q(p)= \frac{\int_{q \in Q} {\rm dist}(p,q) \intd q}{{\rm area}(Q)}, $$
where ${\rm dist}(p,q) = |pq| $ is the Euclidean distance between $p$ and
$q$.
Let $FW_Q$ be the Fermat-Weber center of $Q$, and write
$\mu^*_Q = \min \{\mu_Q(p): p\in \RR^2\} =\mu_Q(FW_Q)$.
Carmi, Har-Peled and Katz~\cite{CHK05} showed that there exists a constant
$c>0$ such that $\mu^*_Q \geq c \cdot \Delta(Q) $ holds for any convex body $Q$,
where $\Delta(Q)$ denotes the diameter of $Q$. The convexity is necessary,
since it is easy to construct nonconvex regions where the average
distance from the Fermat-Weber center is arbitrarily small compared to
the diameter. Of course the opposite inequality $\mu^*_Q \leq c'
\cdot \Delta(Q) $ holds for any body $Q$ (convexity is not required),
since we can trivially take $c'=1$.
Let $c_1$ denote the infimum, and $c_2$ denote the supremum of
$\mu^*_Q/\Delta(Q)$ over all convex bodies $Q$ in the plane.
Carmi, Har-Peled and Katz~\cite{CHK05} conjectured that
$c_1 =\frac{1}{6}$ and $c_2=\frac{1}{3}$. Moreover, they conjectured
that the supremum $c_2$ is attained for a circular disk $D$,
where $\mu^*_D=\frac{1}{3}\cdot \Delta(D)$. They also proved that
$\frac{1}{7} \leq c_1 \leq \frac{1}{6}$.
The inequality $c_1 \leq \frac{1}{6}$ is given by an infinite
sequence of rhombi, $P_\varepsilon$, where one diagonal has some
fixed length, say $2$, and the other diagonal tends to zero; see Fig.~\ref{f1}.
By symmetry, the Fermat-Weber center of a rhombus is its center of
symmetry, and one can verify that $\mu^*_{P_\varepsilon}/\Delta(P_\varepsilon)$
tends to $\frac{1}{6}$. The lower bound for $c_1$ has been recently
further improved by Abu-Affash and Katz from $\frac{1}{7}$ to
$\frac{4}{25}$~\cite{AK08}. Here we establish that $c_1 =\frac{1}{6}$
and thereby confirm the first of the two conjectures of Carmi, Har-Peled and Katz.
\begin{figure} [htb]
\centerline{\epsfxsize=.7\textwidth \epsffile{f1.eps}}
\caption{\small A flat rhombus $P_\varepsilon$, with
$ \lim_{\varepsilon \rightarrow 0} \mu^*_{P_\varepsilon}/\Delta(P_\varepsilon) = \frac{1}{6}$.}
\label{f1}
\end{figure}
Regarding the second conjecture, recently Abu-Affash and Katz
proved that $c_2 \leq \frac{2}{3 \sqrt3} = 0.3849\ldots$. Here we further
improve this bound and bring it closer to the conjectured value of $\frac{1}{3}$.
Finally, we also confirm the upper bound conjecture for centrally
symmetric convex bodies $Q$.
Our main results are summarized in the following two theorems:
\begin{theorem}\label{T1}
For any convex body $Q$ in the plane, we have $\mu^*_Q > \frac{1}{6} \cdot \Delta(Q)$.
\end{theorem}
\begin{theorem}\label{T2}
For any convex body $Q$ in the plane, we have
$$ \mu^*_Q \leq \frac{2(4-\sqrt3)}{13} \cdot \Delta(Q) < 0.3490 \cdot \Delta(Q). $$
Moreover, if $Q$ is centrally symmetric, then $\mu^*_Q \leq
\frac{1}{3} \cdot \Delta(Q)$.
\end{theorem}
\paragraph{Remarks.}
\noindent {\bf 1.}
The average distance from a point $p$ in the plane can be defined analogously
for finite point sets and for rectifiable curves. Observe that for a line segment
$I$ (a one-dimensional convex set), we would have $\mu^*_I/\Delta(I) = \frac{1}{4}$.
It might be interesting to note that while the thin rhombi mentioned
above tend in the limit to a line segment, the value of the
limit $\mu^*_{P_\varepsilon}/\Delta(P_\varepsilon)$ equals $\frac{1}{6}$, not $\frac{1}{4}$.
\medskip
\noindent {\bf 2.}
In some applications, the cost of serving a location $q$
from a facility at point $p$ is ${\rm dist}^\kappa(p,q)$ for some
exponent $\kappa\geq 1$, rather than ${\rm dist}(p,q)$. We can define
$\mu_Q^\kappa(p)=\left(\int_{q\in Q} {\rm dist}^\kappa(p,q)\intd
q\right)/{\rm area}(Q)$
and $\mu_Q^{\kappa *}=\inf \{\mu_Q^\kappa(p): p \in \RR^2\}$, which is
invariant under congruence. The ratio $\mu_Q^{\kappa *} /
\Delta^\kappa(Q)$ is also invariant under similarity. The proof of
Theorem~\ref{T1} carries over for this variant and shows that
$\mu_Q^{\kappa*}/\Delta^\kappa(Q)>\frac{1}{(\kappa+2)2^\kappa}$ for any
convex body $Q$, and $\lim_{\varepsilon\rightarrow 0} \mu_{P_\varepsilon}^{\kappa
*}/2^\kappa
=\frac{1}{(\kappa+2)2^\kappa}$.
For the upper bound, the picture is not so clear: $\mu_Q^*/\Delta(Q)$ is
conjectured to be maximal for the circular disk, however,
there is a $\kappa\geq 1$ such that $\mu_Q^{\kappa *}/\Delta^\kappa(Q)$
cannot be maximal for the disk. In particular, if $D$ is a disk of diameter
2 and $R$ is a convex body of diameter $2$ whose smallest enclosing disk
has diameter more than $2$ (e.g., a regular or a Reuleaux triangle of
diameter $2$), then $\mu_D^{\kappa *} <\mu_{R}^{\kappa *}$,
for a sufficiently large $\kappa>1$.
Let $o$ be an arbitrary point in the plane, and let
$D$ be centered at $o$.
Then $\int_{q\in D}{\rm dist}^\kappa(o,q)\intd q =
\int_0^{2\pi} \int_0^1 r^\kappa \cdot r \intd r \intd \theta =
\frac{2\pi}{\kappa+2}$,
and so $\lim_{\kappa\rightarrow \infty}\mu_D^{\kappa *} \leq
\lim_{\kappa\rightarrow \infty} \frac{2}{\kappa+2}= 0$.
On the other hand, for any region $R'$ lying outside of $D$ and
for any $\kappa\geq 1$, we have $\int_{q\in R'} {\rm dist}^\kappa(o,q)
\intd q\geq {\rm area}(R')>0$. If $R'=R\setminus D$ is the part of $R$ lying
outside $D$, then
$\lim_{\kappa\rightarrow \infty}\mu_{R}^{\kappa *} \geq {\rm
area}(R')/\pi>0$.
\paragraph{Related work.}
Fekete, Mitchell, and Weinbrecht~\cite{FMW00} studied a continuous version
of the problem for polygons with holes, where the distance between two
points is measured by the $L_1$ geodesic distance.
A related question on Fermat-Weber centers in a discrete setting
deals with stars and Steiner stars~\cite{DTX09,FM00}.
\later{
A Steiner star for a set $S$ of $n$ points in $\RR^2$ connects an
arbitrary center point to all points of $S$, while a
star connects a point $p\in S$ to the remaining $n-1$ points of
$S$. All connections are realized by straight line segments.
The maximum ratio between the lengths of the minimum star and the
minimum Steiner star, over all finite point configurations in $\RR^2$,
is called the star Steiner ratio in $\RR^2$, denoted by $\rho$.
Fekete and Meijer~\cite{FM00}, who were the first to study the star Steiner
ratio, proved that $\rho \leq \sqrt{2}$. The current best upper bound
is (about) $1.3631$, see~\cite{DTX09}.
}
The reader can find more information on other variants of the
Fermat-Weber problem in~\cite{DKSW02,W93}.
\section{Lower bound: proof of Theorem~\ref{T1}} \label{sec:T1}
In a nutshell the proof goes as follows.
Given a convex body $Q$, we take its Steiner symmetrization with
respect to a supporting line of a diameter segment $cd$, followed by another
Steiner symmetrization with respect to the perpendicular bisector of
$cd$. The two Steiner symmetrizations preserve the area and the
diameter, and do not increase the average distance from the
corresponding Fermat-Weber centers.
In the final step, we prove that the inequality holds for
a convex body with two orthogonal symmetry axes.
\paragraph{Steiner symmetrization with respect to an axis.}
Steiner symmetrization of a convex figure $Q$ with respect to an axis
(line) $\ell$ consists in replacing $Q$ by a new figure $S(Q,\ell)$ with
symmetry axis $\ell$ by means of the following construction:
Each chord of $Q$ orthogonal to $\ell$ is displaced along its line to
a new position where it is symmetric with respect to $\ell$,
see~\cite[pp.~64]{YB61}. The resulting figure $S(Q,\ell)$ is also
convex, and obviously has the same area as $Q$.
A body $Q$ is $x$-monotone if the intersection of $Q$ with every
vertical line is either empty or is connected (that is, a point
or a line segment).
Every $x$-monotone body $Q$ is bounded by the graphs of some functions
$f: [a,b]\rightarrow \RR$ and $g: [a,b]\rightarrow \RR$ such that
$g(x)\leq f(x)$ for all $x\in [a,b]$. The Steiner symmetrization with
respect to the $x$-axis $\ell_x$ transforms $Q$ into an $x$-monotone body
$S(Q,\ell_x)$ bounded by the functions $\frac{1}{2}(f(x)-g(x))$ and
$\frac{1}{2}(g(x)-f(x))$ for $x\in [a,b]$.
As noted earlier, ${\rm area}(S(Q,\ell_x))={\rm area}(Q)$.
The next two lemmas do not require the convexity of $Q$.
\begin{lemma} \label{L1}
Let $Q$ be an $x$-monotone body in the plane with a diameter parallel or
orthogonal to the $x$-axis, then $\Delta(Q)=\Delta(S(Q,\ell_x))$.
\end{lemma}
\begin{proof}
Let $Q'=S(Q,\ell_x)$. If $Q$ has a diameter parallel to the $x$-axis,
then the diameter is $[(a,c), (b,c)]$, with a value $c\in \RR$,
$g(a)= c= f(a)$ and $g(b)= c= f(b)$. That is, $\Delta(Q)=b-a$.
In this case, the diameter of $Q'$ is at least $b-a$, since both points
$(a,0)$ and $(b,0)$ are in $Q'$. If $Q$ has a diameter orthogonal to
the $x$-axis, then the diameter is
$[(x_0,f(x_0)), (x_0,g(x_0))]$ for some $x_0\in [a,b]$, and
$\Delta(Q)=f(x_0)-g(x_0)$. In this case, the diameter of $Q'$ is at least
$f(x_0)-g(x_0)$, since both points $(x_0,\frac{1}{2}(f(x_0)-g(x_0)))$ and
$(x_0,\frac{1}{2}(g(x_0)-f(x_0)))$ are in $Q'$. Therefore, we have
$\Delta(Q')\geq \Delta(Q)$.
Let $A_1$ and $A_2$ be two points on the boundary of $Q'$ such that
$\Delta(Q')= {\rm dist}(A_1,A_2)$.
Since $Q'$ is symmetric to the $x$-axis, points $A_1$ and $A_2$ cannot
both be on the upper (resp., lower)
boundary of $Q'$. Assume w.l.o.g.\ that $A_1=(x_1,\frac{1}{2}(f(x_1)-g(x_1)))$ and
$A_2=(x_2,\frac{1}{2}(g(x_2)-f(x_2)))$ for some $a\leq x_1,x_2\leq b$.
$$\Delta(Q')= {\rm dist}(A_1,A_2)= \sqrt{(x_2-x_1)^2 +
\left(\frac{f(x_1)+f(x_2)-g(x_1)-g(x_2)}{2}\right)^2}.$$
Now consider the following two point pairs in $Q$. The distance between
$B_1=(x_1,f(x_1))$ and $B_2=(x_2,g(x_2))$ is ${\rm dist}(B_1,B_2)=
\sqrt{(x_2-x_1)^2 + (f(x_1)-g(x_2))^2}$.
Similarly, the distance between $C_1=(x_1,g(x_1))$ and $C_2=(x_2,f(x_2))$ is ${\rm
dist}(C_1,C_2) = \sqrt{(x_2-x_1)^2 + (g(x_1)-f(x_2))^2}$.
Using the inequality between the arithmetic and quadratic means, we have
$$ \left(\frac{f(x_1)+f(x_2)-g(x_1)-g(x_2)}{2}\right)^2 \leq
\frac{(f(x_1)-g(x_2))^2 + (g(x_1)-f(x_2))^2}{2}.$$
This implies that ${\rm dist}(A_1,A_2) \leq \max ( {\rm dist}(B_1,B_2),
{\rm dist}(C_1,C_2))$, and so $\Delta(Q')\leq \Delta(Q)$.
We conclude that $\Delta(Q)=\Delta(S(Q,\ell_x))$.
\end{proof}
\begin{lemma} \label{L2}
If $Q$ is an $x$-monotone body in the plane, then $\mu_Q^*\geq
\mu_{S(Q,\ell_x)}^*$.
\end{lemma}
\begin{proof}
If $(x_0,y_0)$ is the Fermat-Weber center of $Q$, then
$$\mu_Q^*= \frac{\int_a^b \int_{g(x)}^{f(x)} \sqrt{(x-x_0)^2+(y-y_0)^2} \intd y
\intd x}{{\rm area}(Q)}.$$
Observe that $\int_{g(x)}^{f(x)} \sqrt{(x-x_0)^2+(y-y_0)^2} \intd y$ is the
integral of the distances of the points in a line segment of length
$f(x)-g(x)$ from a point at distance $|x-x_0|$ from the supporting
line of the segment. This integral is minimal if the point is on the
orthogonal bisector of the segment. That is, we have
\begin{eqnarray}
\int_{g(x)}^{f(x)} \sqrt{(x-x_0)^2+(y-y_0)^2} \intd y
&\geq& \int_{g(x)}^{f(x)}
\sqrt{(x-x_0)^2+\left(y-\frac{f(x)-g(x)}{2}\right)^2} \ dy \nonumber\\
&=& \int_{\frac{1}{2}(g(x)-f(x))}^{\frac{1}{2}(f(x)-g(x))}
\sqrt{(x-x_0)^2+y^2} \intd y.\nonumber
\end{eqnarray}
Therefore, we conclude that
\begin{eqnarray}
\mu_Q^*
& = & \frac{\int_a^b \int_{g(x)}^{f(x)} \sqrt{(x-x_0)^2+(y-y_0)^2} \intd y
\intd x}{{\rm area}(Q)}\nonumber\\
&\geq& \frac{\int_a^b \int_{\frac{1}{2}(g(x)-f(x))}^{\frac{1}{2}(f(x)-g(x))}
\sqrt{(x-x_0)^2+y^2} \intd y \intd x}{{\rm area}(S(Q,x))} =
\mu_{S(Q,\ell_x)}((x_0,0))\geq \mu_{S(Q,\ell_x)}^*.\nonumber
\end{eqnarray}
\end{proof}
\paragraph{Triangles.} We next consider right triangles of a special
kind, lying in the first quadrant, and show that the average distance from
the origin to their points is larger than $\frac{1}{3}$.
\begin{lemma} \label{L3}
Let $T$ a right triangle in the first quadrant based on the $x$-axis,
with vertices $(a,0)$, $(a,b)$, and $(1,0)$, where $0 \leq a < 1$, and $b>0$.
Then $\mu_T(o) > \frac{1}{3}$.
\end{lemma}
\begin{proof}
We use the simple fact that the $x$-coordinate of a point is a lower
bound to the distance from the origin.
\begin{eqnarray*}
\mu_T(o) &=&
\frac{\int_a^1 (\int_0^{b(1-x)/(1-a)} \sqrt{x^2 + y^2} \intd y) \intd x}
{b(1-a)/2} >
\frac{\int_a^1 (\int_0^{b(1-x)/(1-a)} x \intd y) \intd x} {b(1-a)/2} \\
&=&
\frac{\frac{b}{1-a} \int_a^1 x(1-x) \intd x} {b(1-a)/2} =
\frac{2}{(1-a)^2}\left(\frac{x^2}{2} - \frac{x^3}{3}\right) \Big{|}_a^1\\
&=& \frac{2}{(1-a)^2} \cdot \frac{(2a^3-3a^2+1)}{6} =
\frac{2}{(1-a)^2} \cdot \frac{(1-a)(1+a-2a^2)}{6} \\
&=& \frac{1}{(1-a)} \cdot \frac{(1+a-2a^2)}{3} \geq \frac{1}{3}.
\end{eqnarray*}
The last inequality in the chain follows from $0 \leq a<1$.
The inequality in the lemma is strict, since $\sqrt{x^2 + y^2} > x$ for all points
above the $x$-axis.
\end{proof}
\begin{corollary} \label{C1}
Let $P$ be any rhombus. Then $\mu^*_P > \frac{1}{6} \cdot \Delta(P)$.
\end{corollary}
\begin{proof}
Without loss of generality, we may assume that $P$ is symmetric with
respect to both the $x$-axis and the $y$-axis. Let us denote the vertices of $P$ by
$(-1,0)$, $(1,0)$, $(0,-b)$, and $(0,b)$, where $b \leq 1$. We have $\Delta(P)=2$.
By symmetry, $\mu^*_P$ equals the average distance between the origin $(0,0)$ and
the points in one of the four congruent right triangles forming $P$.
Consider the triangle $T$ in the first quadrant.
By Lemma~\ref{L3} (with $a=0$), we have $\mu^*_P = \mu_T(o) > \frac{1}{3}$.
Since $\Delta(P)=2$, we have $\mu^*_P > \frac{1}{6} \cdot \Delta(P)$, as desired.
\end{proof}
\begin{lemma} \label{L4}
Let $T$ be a triangle in the first quadrant with a vertical side
on the line $x=a$, where $0 \leq a <1$, and a third vertex at
$(1,0)$. Then $\mu_T(o) > \frac{1}{3}$.
\end{lemma}
\begin{proof}
Refer to Fig.~\ref{f2}(ii). Let $U$ be a right triangle obtained from
$T$ by translating each vertical chord of $T$ down until its lower
endpoint is on the $x$-axis. Note that ${\rm area}(T)={\rm area}(U)$.
Observe also that the average distance from the origin decreases in this
transformation, namely $ \mu_T(o) \geq \mu_U(o)$.
By Lemma~\ref{L3}, we have $\mu_U(o) > \frac{1}{3}$, and so
$\mu_T(o) > \frac{1}{3}$, as desired.
\end{proof}
\smallskip
We now have all necessary ingredients to prove Theorem~\ref{T1}.
\paragraph {Proof of Theorem~\ref{T1}.}
Refer to Fig.~\ref{f2}.
Let $Q$ be a convex body in the plane, and let $c,d\in Q$ be two
points at $\Delta(Q)$ distance apart. We may assume that $c=(-1,0)$
and $d=(1,0)$, by a similarity transformation if necessary, so that
$\Delta(Q)=2$ (the ratio $\mu^*_Q/\Delta(Q)$ is invariant under similarities).
Apply a Steiner symmetrization with respect to the $x$-axis, and then a second
Steiner symmetrization with respect to the $y$-axis.
The resulting body $Q'=S(S(Q,\ell_x), \ell_y)$ is convex, and it is
symmetric with respect to both coordinate axes.
We have $\Delta(Q')=\Delta(Q)=2$ by Lemma~\ref{L1}, and in fact $c,d\in Q'$.
We also have $\mu^*_{Q'}\leq \mu^*_Q$ by Lemma~\ref{L2}.
\begin{figure} [b]
\centerline{\epsfxsize=.97\textwidth \epsffile{f2.eps}}
\caption{\small (i) The subdivision of $Q_1$ for $n=3$. Here
$o=(0,0)$, $q_1=b=(0,h)$, $q_4=r$, $d=(1,0)$.
(ii) Transformation in the proof of Lemma~\ref{L4}.}
\label{f2}
\end{figure}
Let $Q_1$ be the part of $Q'$ lying in the first quadrant:
$Q_1 =\{(x,y) \in Q': x,y \geq 0\}$.
By symmetry, $FW_{Q'}=o$ and we have $\mu^*_{Q'}=\mu_{Q'}(o)=\mu_{Q_1}(o)$.
Let $\gamma$ be the portion of the boundary of $Q'$ lying in the first quadrant,
between points $b=(0,h)$, with $0<h\leq 1$, and $d=(1,0)$. For any two
points $p,q\in \gamma$ along $\gamma$, denote by $\gamma(p,q)$ the
portion of $\gamma$ between $p$ and $q$.
Let $r$ be the intersection point of $\gamma$ and the vertical line $x=\frac{1}{3}$.
For a positive integer $n$, subdivide $Q_1$ into at most $2n+2$ pieces as follows.
Choose $n+1$ points $b=q_1,q_2\ldots , q_{n+1}=r$ along $\gamma(b,r)$
such that $q_i$ is the intersection of $\gamma$ and the vertical line $x=(i-1)/3n$.
Connect each of the $n+1$ points to $d$ by a straight line segment. These segments
subdivide $Q_1$ into $n+2$ pieces: the right triangle $T_0=\Delta bod$; a
convex body $Q_0$ bounded by $rd$ and $\gamma(r,d)$; and $n$ curvilinear triangles
$\Delta q_idq_{i+1}$ for $i=1,2,\ldots , n$. For simplicity, we assume
that neither $Q_0$, nor any of the curvilinear triangles are degenerate;
otherwise they can be safely ignored (they do not contribute
to the value of $\mu^*_{Q'}$). Subdivide each curvilinear triangle
$\Delta q_idq_{i+1}$ along the vertical line through $q_{i+1}$ into
a small curvilinear triangle $S_i$ on the left and a triangle $T_i$ incident
to point $d$ on the right. The resulting subdivision has $2n+2$ pieces,
under the nondegeneracy assumption.
By Lemma~\ref{L3}, we have $\mu_{T_0}(o)>\frac{1}{3}$. Observe that the difference
$\mu_{T_0}(o)-\frac{1}{3}$ does not depend on $n$, and let $\delta
=\mu_{T_0}(o)-\frac{1}{3}$.
By Lemma~\ref{L4}, we also have $\mu_{T_i}(o)>\frac{1}{3}$, for
each $i=1,2,\ldots ,n$.
Since every point in $Q_0$ is at distance at least $\frac{1}{3}$ from the origin,
we also have $\mu_{Q_0}(o)\geq \frac{1}{3}$.
For the $n$ curvilinear triangles $S_i$, $i=1,2,\ldots ,n$, we use the
trivial lower bound $\mu_{S_i}(o)\geq 0$. We now show that their total area
$s_n=\sum_{i=1}^n {\rm area}(S_i)$ tends to 0 if $n$ goes to infinity.
Recall that the $y$-coordinates of the points $q_i$ are at most 1, and
their $x$-coordinates are at most $\frac{1}{3}$. This implies that the slope
of every line $q_id$, $i=1,2,\ldots , n+1$,
is in the interval $[-3/2, 0]$. Therefore, $S_i$ is contained in a right triangle
bounded by a horizontal line through $q_i$, a vertical line through $q_{i+1}$, and
the line $q_id$. The area of this triangle is at most $\frac{1}{2}(\frac{1}{3n}\cdot
(\frac{3}{2}\cdot \frac{1}{3n})) = 1/(12n^2)$.
That is, $s_n =\sum_{i=1}^n {\rm area}(S_i) \leq 1/(12n)$.
In particular, $s_n \leq \delta \cdot {\rm area}(T_0)$ for a sufficiently large $n$.
Then we can write
\begin{eqnarray*}
\mu_{Q_1}(o) &=& \frac{\int_{p\in Q_1}{\rm dist}(o,p) \intd p}{{\rm area}(Q_1)}
\geq \frac{ \mu_{Q_0}(o) \cdot {\rm area}(Q_0)+
\sum_{i=0}^n \mu_{T_i}(o) \cdot {\rm area}(T_i)}{{\rm area}(Q_1)} \\
&\geq& \frac{\frac{1}{3}({\rm area}(Q_1)-s_n)+\delta \cdot {\rm area}(T_0)}
{{\rm area}(Q_1)}
\geq \frac{1}{3}+ \frac{2\delta \cdot {\rm area}(T_0)} {3 \cdot {\rm area}(Q_1)}
>\frac{1}{3}.
\end{eqnarray*}
This concludes the proof of Theorem~\ref{T1}.
\hfill $\Box$ \linebreak \smallskip
\medskip
\noindent {\bf Remark.} A finite triangulation, followed by
taking the limit suffices to prove the slightly weaker, non-strict
inequality: $\mu^*_Q \geq \frac{1}{6} \cdot \Delta(Q)$.
\section{Upper bounds: proof of Theorem~\ref{T2}} \label{sec:T2}
Let $Q$ be a planar convex body and let $D=\Delta(Q)$.
Let $\partial Q$ denote the boundary
of $Q$, and let ${\rm int}(Q)$ denote the interior of $Q$.
Let $\Omega$ be the smallest disk enclosing $Q$, and let $o$ and
$R$ be the center and respectively the radius of $\Omega$.
Write $a= \frac{2(4-\sqrt3)}{13}$.
By the convexity of $Q$, $o \in Q$, as observed in~\cite{AK08}.
Moreover, Abu-Affash and Katz~\cite{AK08} have shown that the average
distance from $o$ to the points in $Q$ satisfies
$$ \mu_Q(o) \leq \frac{2}{3 \sqrt3} \cdot \Delta(Q) < 0.3850 \cdot \Delta(Q). $$
Here we further refine their analysis and derive a better upper bound
on the average distance from $o$ to the points in $Q$:
$$ \mu_Q(o) \leq \frac{2(4-\sqrt3)}{13} \cdot \Delta(Q) < 0.3490 \cdot \Delta(Q). $$
Since the average distance from the Fermat-Weber center of $Q$ is not
larger than that from $o$, we immediately get the same upper bound on $c_2$.
We need the next simple lemma established in~\cite{AK08}. Its
proof follows from the definition of average distance.
\begin{lemma} {\rm~\cite{AK08}.} \label{L-AK}
Let $Q_1$, $Q_2$ be two (not necessarily convex) disjoint bodies in
the plane, and $p$ be a point in the plane. Then
$\mu_{(Q_1 \cup Q_2)}(p) \leq \max (\mu_{Q_1}(p), \mu_{Q_2}(p))$.
\end{lemma}
By induction, Lemma~\ref{L-AK} yields:
\begin{lemma} \label{L5}
Let $Q_1, Q_2, \ldots, Q_n$ be $n$ (not necessarily convex) pairwise
disjoint bodies in the plane, and $p$ be a point in the plane. Then
$$ \mu_{(Q_1 \cup \ldots \cup Q_n)}(p) \leq
\max (\mu_{Q_1}(p), \ldots \mu_{Q_n}(p)). $$
\end{lemma}
We also need the following classical result of Jung~\cite{J10};
see also~\cite{HD64}.
\begin{theorem} {\rm (Jung~\cite{J10}).} \label{T-J}
Let $S$ be a set of diameter $\Delta(S)$ in the plane. Then
$S$ is contained in a circle of radius $\frac{1}{\sqrt3} \cdot \Delta(S)$.
\end{theorem}
By Theorem~\ref{T-J} we have
\begin{equation}\label{eq:RD}
\frac12 D \le R \le \frac1{\sqrt3} D.
\end{equation}
Observe that the average distance from the center
of a circular sector of radius $r$ and center angle $\alpha$ to the
points in the sector is
\begin{equation} \label{E1}
\frac{\int_0^r \alpha x^2 \intd x}{\int_0^r \alpha x \intd x}=
\frac{\alpha r^3/3}{\alpha r^2/2} = \frac{2r}{3}.
\end{equation}
\paragraph {Proof of Theorem~\ref{T2}.}
If $o \in \partial Q$ then $Q$ is contained in a
halfdisk $\Theta$ of $\Omega$, of the same diameter $D$, with $o$ as
the midpoint of this diameter. Then by \eqref{E1}, it follows that
$ \mu_Q(o) \leq \frac{1}{3} \cdot D$, as required.
We can therefore assume that $o \in {\rm int}(Q)$.
Let $\varepsilon>0$ be sufficiently small.
For a large positive integer $n$, subdivide $\Omega$
into $n$ congruent circular double sectors (wedges) $W_1,\ldots,W_n$,
symmetric about $o$ (the center of $\Omega$), where each sector
subtends an angle $\alpha=\pi/n$. Consider a double sector $W_i =U_i \cup V_i$,
where $U_i$ and $V_i$ are circular sectors of $\Omega$.
Let $X_i \subseteq U_i$, and $Y_i \subseteq V_i$
be two minimal circular sectors centered at $o$ and
containing $U_i \cap Q$, and $V_i \cap Q$, respectively:
$ U_i \cap Q \subseteq X_i$, and $ V_i \cap Q \subseteq Y_i$.
Let $x_i$ and $y_i$ be the radii of $X_i$ and $Y_i$, respectively.
Let $X'_i \subseteq X_i$, and $Y'_i \subseteq Y_i$ be two circular
subsectors of radii $(1-\varepsilon)x_i$ and $(1-\varepsilon)y_i$, respectively.
Since $o \in {\rm int}(Q)$, we can select $n=n(Q,\varepsilon)$ large enough, so
that for each $1 \leq i \leq n$, the subsectors $X'_i$ and $Y'_i$
are nonempty and entirely contained in $Q$. That is, for every $i$, we have
\begin{equation} \label{E2}
X'_i \cup Y'_i \subseteq W_i \cap Q \subseteq X_i \cup Y_i.
\end{equation}
It is enough to show that for any double sector $W=W_i$, we have
$$ \lim_{\varepsilon \to 0} \mu_{(W \cap Q)}(o) \leq a D, $$
since then, Lemma~\ref{L5} (with $W_i$ being the $n$ pairwise disjoint
regions) will imply that $ \mu_Q(o) \leq a D$, concluding the proof of
Theorem~\ref{T2}. For simplicity, write $x=x_i$, and $y=y_i$.
Obviously the diameter of $W \cap Q$ is at most $D$, hence $x+y \leq D$.
We can assume w.l.o.g.\ that $y \leq x$, so by Theorem~\ref{T-J}
we also have $x \leq \frac{1}{\sqrt3} \cdot D $. Hence so far, our
constraints are:
\begin{equation} \label{E3}
0< y \leq x \leq \frac{1}{\sqrt3} \cdot D
\hspace{1cm}\mbox{\rm and}\hspace{1cm} x+y \leq D.
\end{equation}
By the minimality of the disk $\Omega$, the convex body $Q$
either contains three points $q_1,q_2,q_3$ on the boundary of $\Omega$
such that the triangle $q_1q_2q_3$ contains the disk center $o$ in the interior,
or contains two points $q_1,q_2$ on the boundary of $\Omega$
such that the segment $q_1q_2$ goes through the disk center $o$.
In the latter case, the segment $q_1q_2$ can be viewed as a degenerate triangle
$q_1q_2q_3$ with two coinciding vertices $q_2$ and $q_3$.
Let $r$ be the radius of the largest disk centered at $o$
that is contained in the convex body $Q$.
Then $r$ is at least the distance from $o$ to the longest side of the triangle
$q_1q_2q_3$, say $q_1q_2$.
Since $|q_1q_2| \le D$, $|o q_1| = |o q_2| = R$,
we have
$$
r \ge \sqrt{R^2 - D^2/4}.
$$
Then the constraints in~\eqref{E3} can be expanded to the following:
\begin{equation}\label{eq:constraints-new}
\sqrt{R^2 - D^2/4} \le y \le x \le R \le D/\sqrt3
\quad\textup{and}\quad
x + y \le D.
\end{equation}
By the definition of average distance, we can write
\begin{eqnarray} \label{E4}
\mu_{(W \cap Q)}(o) &=&
\frac{\int_{p\in (W \cap Q)}{\rm dist}(o,p) \intd p}{{\rm area}(W \cap Q)}\nonumber\\
&\leq& \dfrac{\alpha \cdot \frac{x^2}{2} \cdot \frac{2x}{3} +
\alpha \cdot \frac{y^2}{2} \cdot \frac{2y}{3} }
{\alpha (1-\varepsilon)^2 \cdot \left(\frac{x^2}{2} + \frac{y^2}{2}\right)}
= \frac{2}{3} \cdot \frac{x^3+y^3}{(1-\varepsilon)^2 \cdot (x^2+y^2)}.
\end{eqnarray}
Let
\begin{equation}\label{eq:f}
f(x,y)= \frac{2}{3} \cdot \frac{x^3+y^3}{x^2+y^2},
\textrm{ \ and \ }
f_1(x,y,\varepsilon)= \frac{2}{3} \cdot \frac{x^3+y^3}{(1-\varepsilon)^2 \cdot (x^2+y^2)}.
\end{equation}
Clearly for any feasible pair $(x,y)$, we have
$$
\lim_{\varepsilon \to 0} f_1(x,y,\varepsilon) = f(x,y).
$$
It remains to maximize $f(x,y)$ subject to the constraints in
\eqref{eq:constraints-new}. We will show that under these constraints,
\begin{equation}\label{E10}
f(x,y) \leq \frac{2(4-\sqrt3)}{13} \cdot D.
\end{equation}
Then
$$
\lim_{\varepsilon \to 0} \mu_{(W \cap Q)}(o) \leq
\lim_{\varepsilon \to 0} f_1(x,y,\varepsilon) = f(x,y)
\leq \frac{2(4-\sqrt3)}{13} \cdot D,
$$
as required.
We next verify the upper bound in~\eqref{E10}.
Throughout our analysis,
we may assume that $D$ is a fixed constant and
$x$, $y$, and $R$ are variable parameters.
Substituting $z = y/x$ in~\eqref{eq:f}, we have
$$
f(x, y) = g(x, z) = \frac{2x}3 \cdot \frac{1 + z^3}{1 + z^2}.
$$
Then, taking the partial derivative of $g(x,z)$ with respect to $z$, we have
\begin{align*}
\frac{\partial}{\partial z} g(x, z) &= \frac{2x}3 \cdot
\left( \frac{3 z^2}{1 + z^2} - \frac{1 + z^3}{(1 + z^2)^2}
2z\right) \\
&= \frac{2x}3 \cdot \frac{3 z^2 (1 + z^2) - (1 + z^3) 2z}{(1 + z^2)^2}
= \frac{2x}3 \cdot \frac{z (z^3 + 3 z - 2)}{(1 + z^2)^2}.
\end{align*}
The cubic equation $z^3 + 3 z - 2 = 0$ has exactly one real root
$z_0 = (\sqrt2 + 1)^{1/3} - (\sqrt2 - 1)^{1/3} = 0.596\ldots.$
Thus for a fixed $x$, the function $g(x, z)$ is strictly decreasing for
$0 \le z \le z_0$ and is strictly increasing for $z_0 \le z \le 1$.
Therefore,
by the upper bound that
$x + y \le D$
and the lower bound that
$\sqrt{R^2 - D^2/4} \le r \le y$
in~\eqref{eq:constraints-new},
the function $f(x, y)$ is maximized when $y$ takes one of the following
two extreme values:
$$
y_1 = \sqrt{R^2 - D^2/4}
\quad\textup{and}\quad
y_2 = D - x.
$$
By the inequality that $x \le R \le D/\sqrt3$ in~\eqref{eq:constraints-new},
it follows that
$x + y_1 \le R + \sqrt{R^2 - D^2/4} \le D/\sqrt3 + D/\sqrt{12} < D$.
Since
$x + y_2 = D$,
we have $y_1 < y_2$.
\paragraph{Case 1.}
We first consider the easy case that $y = y_2$. Then $x + y = D$,
and we have
$$
f(x, y) = \frac23 \cdot \frac{x^3 + y^3}{x^2 + y^2}
= \frac23 \cdot \frac{(x+y)^3 - 3(x+y)xy}{(x+y)^2 - 2xy}
= \frac23 \cdot \frac{D^3 - 3Dxy}{D^2 - 2xy}.
$$
Substituting $w = xy$,
we tranform the function $f(x,y)$ to a function $h_1(w)$:
$$
f(x, y) = h_1(w) = \frac23 \cdot \frac{3Dw - D^3}{2w - D^2}.
$$
The function $h_1(w)$ is decreasing in $w$ because
\begin{align*}
\frac{\mathrm{d}}{\mathrm{d} w} h_1(w) &= \frac23 \cdot
\left( \frac{3D}{2w - D^2} - \frac{2(3Dw - D^3)}{(2w - D^2)^2}
\right) \\
&= \frac23 \cdot \frac{3D(2w - D^2) - 2(3Dw - D^3)}{(2w - D^2)^2}
= \frac23 \cdot \frac{-D^3}{(2w - D^2)^2} \le 0.
\end{align*}
Thus $f(x,y)$ is maximized when $xy$ is minimized.
With the sum $x + y$ fixed at $D$,
and under the constraint that $x \le R \le D/\sqrt3$
in~\eqref{eq:constraints-new},
the product $xy$ is minimized when
$x = \frac1{\sqrt3} D$ and $y = \left(1 - \frac1{\sqrt3}\right) D$.
Thus we have
\begin{equation}\label{eq:case1}
f(x, y) \le \frac23 \cdot
\frac{\left(\frac1{\sqrt3}\right)^3 + \left(1 - \frac1{\sqrt3}\right)^3}
{\left(\frac1{\sqrt3}\right)^2 + \left(1 - \frac1{\sqrt3}\right)^2} D
= \frac{2(4 - \sqrt3)}{13} D
= 0.3489\ldots D.
\end{equation}
\paragraph{Case 2.}
We next consider the case\footnote{This case, when $x+y <D$,
has been mistakenly overlooked in the proof given in~\cite{DT09}.}
that $y = y_1$.
With $y$ fixed,
the function $f(x,y)$ is maximized when $x$ is as large as possible because
\begin{align*}
\frac{\partial}{\partial x} f(x, y) &= \frac23 \cdot
\left( \frac{3x^2}{x^2 + y^2} - \frac{x^3 + y^3}{(x^2 + y^2)^2}2x \right)
\\
&= \frac23 \cdot \frac{3x^2(x^2 + y^2) - (x^3 + y^3)2x}{(x^2 + y^2)^2}
\\
&= \frac23 \cdot \frac{x(x^3 + 3xy^2 - 2y^3)}{(x^2 + y^2)^2}
\\
&\ge \frac23 \cdot \frac{x(y^3 + 3y^3 - 2y^3)}{(x^2 + y^2)^2} \ge 0.
\end{align*}
Thus for $y= \sqrt{R^2 - D^2/4}$ and
under the constraint that $x \le R$ in~\eqref{eq:constraints-new},
the function $f(x, y)$ is maximized when
$x = R$ and
$y = \sqrt{R^2 - D^2/4} = \sqrt{x^2 - D^2/4}$.
It follows that
$$
\frac{\mathrm{d} x}{\mathrm{d} R} = 1
\qquad\textup{and}\qquad
\frac{\mathrm{d} y}{\mathrm{d} R} = \frac{\mathrm{d} \sqrt{x^2 - D^2/4}}{\mathrm{d} R}
= \frac{x}{\sqrt{x^2 - D^2/4}} = x/y.
$$
Let $h_2(R) = f(R, \sqrt{R^2 - D^2/4})$.
We next show that $h_2(R)$ is increasing in $R$.
Taking the derivative, we have
\begin{align*}
\frac{\mathrm{d}}{\mathrm{d} R} h_2(R) &= \frac23 \cdot
\left( \frac{3x^2 \frac{\mathrm{d} x}{\mathrm{d} R} + 3y^2 \frac{\mathrm{d} y}{\mathrm{d} R}}{x^2 + y^2}
- \frac{x^3 + y^3}{(x^2 + y^2)^2}
\left( 2x \frac{\mathrm{d} x}{\mathrm{d} R} + 2y \frac{\mathrm{d} y}{\mathrm{d} R} \right)
\right)
\\
&= \frac23 \cdot
\left( \frac{3x^2 + 3y^2 (x/y)}{x^2 + y^2}
- \frac{x^3 + y^3}{(x^2 + y^2)^2}(2x + 2y(x/y))
\right)
\\
&= \frac23 \cdot
\frac{(3x^2 + 3xy)(x^2 + y^2) - (x^3 + y^3)(2x + 2x)}{(x^2 + y^2)^2}
\\
&= \frac23 \cdot
\frac{(3x^4 + 3x^2y^2 + 3x^3y + 3xy^3) - (4x^4 + 4xy^3)}{(x^2 + y^2)^2}
\\
&= \frac23 \cdot
\frac{(x^4 + 3x^2y^2 + 3x^3y + xy^3) - (2x^4 + 2xy^3)}{(x^2 + y^2)^2}
\\
&= \frac23 \cdot
\frac{x^4}{(x^2 + y^2)^2} \cdot \big( (1+y/x)^3 - 2 - 2 (y/x)^3 \big).
\end{align*}
Substituting $z = y/x$,
we simplify the last factor
$(1+y/x)^3 - 2 - 2 (y/x)^3$
in the resulting expression above to
$$
h_3(z) = (1+z)^3 - 2 - 2 z^3.
$$
To show that $\frac{\mathrm{d}}{\mathrm{d} R} h_2(R) > 0$,
it remains to show that $h_3(z) > 0$.
For $0 \le z \le 1$,
the function $h_3(z)$ is increasing in $z$ because
$$
\frac{\mathrm{d}}{\mathrm{d} z} h_3(z) = 3(1+z)^2 - 6 z^2 = -3(1-z)^2 + 6
\geq 6-3 > 0.
$$
Recall that $x \ge y$.
If $R \le \frac{3(4 - \sqrt3)}{13} D$,
then we would easily have
$$
f(x, y) = \frac23 \cdot \frac{x^3 + y^3}{x^2 + y^2}
\le \frac23 \cdot \frac{x^3}{x^2}
= \frac23 x \le \frac 23 R \le \frac{2(4 - \sqrt3)}{13} D,
$$
which matches the upper bound in case~1.
Now suppose that $R > \frac{3(4 - \sqrt3)}{13} D$.
Then
$$ D/R < \frac{13}{3(4 - \sqrt3)}
{\rm \ and \ }
z = y/x = \sqrt{1 - (D/R)^2/4} >
\sqrt{1 - \left(\frac{13}{3(4 - \sqrt3)}\right)^2 \bigg{/} 4 }
= 0.2955\ldots. $$
It follows that
$$ h_3(z) > h_3\left(\sqrt{1 - \left(\frac{13}{3(4 - \sqrt3)}\right)^2 \bigg{/} 4 }\right)
= 0.1226\ldots > 0, $$
hence
$$ \frac{\mathrm{d}}{\mathrm{d} R} h_2(R) > 0. $$
We have shown that the function $h_2(R)$ is increasing in $R$.
Then,
under the constraint that $R \le D / \sqrt3$ in~\eqref{eq:constraints-new},
$h_2(R)$ is maximized when $R = \frac1{\sqrt3} D$.
Correspondingly,
$f(x,y)$ is maximized when
$x = \frac1{\sqrt3} D$ and $y = \frac1{\sqrt{12}} D$.
Thus
\begin{equation}\label{eq:case2}
f(x, y) \le \frac23 \cdot
\frac{\left(\frac1{\sqrt3}\right)^3 + \left(\frac1{\sqrt{12}}\right)^3}
{\left(\frac1{\sqrt3}\right)^2 + \left(\frac1{\sqrt{12}}\right)^2} D
= \frac{\sqrt3}5 D
= 0.3464\ldots D,
\end{equation}
which is (slightly) smaller than the upper bound obtained in case~1.
This proves the upper bound in \eqref{E10}.
\paragraph {Centrally symmetric body.}
Assume now that $Q$ is centrally symmetric with respect to a point $q$.
We repeat the same ``double sector'' argument. It is enough to observe
that: (i) the center of $\Omega$ coincides with $q$, that is, $o=q$; and (ii) $x=y
\leq \frac{1}{2} \cdot D$ for any double sector $W$.
By \eqref{E4}, the average distance calculation yields now
$$ \mu_{(W \cap Q)}(o) \leq \frac{2x^3}{3(1-\varepsilon)^2 \cdot x^2}
= \frac{2x}{3(1-\varepsilon)^2} \leq \frac{D}{3(1-\varepsilon)^2}, $$
and by taking the limit when $\varepsilon$ tends to zero, we obtain
$$ \mu_Q(o) \leq \frac{D}{3}, $$
as required. The proof of Theorem~\ref{T2} is now complete.
\hfill $\Box$ \linebreak \smallskip
\section{Applications} \label{sec:app}
\noindent {\bf 1.}
Carmi, Har-Peled and Katz~\cite{CHK05} showed that given a convex
polygon $Q$ with $n$ vertices, and a parameter $\varepsilon>0$, one can
compute an $\varepsilon$-approximate Fermat-Weber center $q \in Q$ in
$O(n+1/\varepsilon^4)$ time such that $\mu_Q(q) \leq (1+\varepsilon) \mu^*_Q$.
Abu-Affash and Katz~\cite{AK08} gave a simple $O(n)$-time algorithm for
computing the center $q$ of the smallest disk enclosing $Q$, and showed that
$q$ approximates the Fermat-Weber center of $Q$, with
$\mu_Q(q) \leq \frac{25}{6 \sqrt3} \mu^*_Q$.
Our Theorems~\ref{T1} and~\ref{T2}, combined with their analysis,
improves the approximation ratio to about $2.09$:
$$ \mu_Q(q) \leq \frac{12(4-\sqrt3)}{13} \mu^*_Q. $$
\smallskip
\noindent {\bf 2.}
The value of the constant $c_1$ (i.e., the infimum of $\mu^*_Q/\Delta(Q)$
over all convex bodies $Q$ in the plane) plays a key role in the following
load balancing problem introduced by Aronov, Carmi and Katz~\cite{ACK06}.
We are given a convex body $D$ and $m$ points $p_1,p_2,\ldots , p_m$
representing {\em facilities} in the interior of $D$. Subdivide $D$ into $m$
convex regions, $R_1,R_2,\ldots ,R_m$, of equal area such that $\sum_{i=1}^m
\mu_{p_i}(R_i)$ is minimal. Here $\mu_{p_i}(R_i)$ is the {\em cost} associated
with facility $p_i$, which may be interpreted as the average travel time from the
facility to any location in its designated region, each of which has the same area.
One of the main results in~\cite{ACK06} is a $(8+\sqrt{2\pi})$-factor approximation
in the case that $D$ is an $n_1\times n_2$ rectangle for some integers $n_1,n_2\in \NN$.
This basic approximation bound is then used for several other cases, e.g.,
subdividing a convex fat domain $D$ into $m$ convex regions $R_i$.
By substituting $c_1=\frac{1}{6}$ (Theorem~\ref{T1}) into the analysis
in~\cite{ACK06}, the upper bound for the approximation ratio improves from
$8+\sqrt{2\pi}\approx 10.5067$ to $7+\sqrt{2\pi}\approx 9.5067$.
It can be further improved by optimizing another parameter
used in their calculation. Let $S$ be a unit square and let $s \in S$
be an arbitrary point in the square. Aronov et al.~\cite{ACK06} used the
upper bound $\mu_S(s)\leq \frac{2}{3}\sqrt{2}\approx 0.9429$. It is
clear that $\max_{s\in S}\mu_S(s)$ is attained if $s$ is a vertex of $S$.
The average distance of $S$ from such a vertex, say $v$, is $\mu_S(v)
= \frac{1}{3}\left(\sqrt{2}+\ln(1+\sqrt{2})\right)\approx 0.7652,$
and so $\mu_S(s)\leq \frac{1}{3}\left(\sqrt{2}+\ln(1+\sqrt{2})\right)$,
for any $s\in S$. With these improvements, the upper
bound on the approximation ratio becomes
$7+\frac{\sqrt{\pi}}{2}\left(\sqrt{2}+\ln(1+\sqrt{2})\right)\approx 9.0344$.
\paragraph{Acknowledgment.} We are grateful to Alex Rand for
stimulating discussions.
\vspace{-.6\baselineskip}
|
\section{Introduction}
It is a classical result by D. Alekseevskii \cite{a} that the group of
conformal automorphisms of
a Riemannian manifold either fixes a conformally equivalent metric or
the manifold is conformally flat. A gap in his proof, found by
R.J. Zimmer and K.R. Gutschera in 1992, was filled by J. Ferrand \cite{f}.
From an infinitesimal point
of view, if a conformal manifold admits a complete
and {\em essential} conformal vector field (whose global
flow acts by conformal transformations, but not by isometries with
respect to any compatible metric), then it is conformally flat,
\cite{kr}. More recently, Ch. Frances proved a local version of this result
\cite{fr}, see also \cite{fm} and \cite {l} for the pseudo-Riemannian setting.
For a given conformal vector field $\xi$ (whose flow is
not globally defined in general), there might exist local metrics in
the conformal class preserved by the local flow of
$\xi$. The union of the definition domains of these $\xi$-invariant
local metrics is the open set of {\em non-essential} points of
the conformal vector field $\xi$. This motivates the following:
\begin{ede} Let $(M,c)$ be a conformal manifold and let $\xi$ be a conformal
vector field on $M$. A point $x\in M$ is called {\em essential} for
$\xi$ if there is no local metric in $c$ preserved by the
local flow of $\xi$ around $x$.
\end{ede}
If $\xi$ does not vanish at some point $p$, it is easy to see that $p$
is not essential. Indeed, $\xi$ is rectifiable in a neighborhood $U$ of
$p$, so there exists a system of coordinates $(x_1,\ldots,x_n)$ near
$p$ such that $x_i(p)=0$ for all $i$ and $\xi=\partial/\partial x_1$. Take any
metric in $c$, restrict it to the hypersurface
$U\cap \{x_1=0\}$ and extend it to $U$ as to be constant with respect
to $x_1$. Then $\xi$ preserves the new metric, which moreover belongs
to the conformal class $c$ because $\xi$ is conformal.
In this note, we prove
\begin{ath}\label{ess} The essential set of a conformal vector field $\xi$ on a
Riemannian manifold $M$ of dimension $n\ge 2$ consists of isolated
zeros of $\xi$.
\end{ath}
Note that this is equivalent for $n=2$ to the fact that the
zeros of a non-constant holomorphic function are isolated.
The theorem above may thus be seen as an extension to higher dimensions of this
classical fact.
For the proof, we focus on the zero set of $\xi$ and show that if a
zero point $x$ of $\xi$ is not isolated then an algebraic condition
(Theorem \ref{cap}) is satisfied by the derivative of $\xi$ at $x$, which
implies, via a result by Beig \cite{beig} and Capocci \cite{C},
that $\xi$ is Killing with respect to a local metric around $x$.
As an application, we show in Theorem \ref{umb} that every connected
component of the zero set of $\xi$ is a totally umbilical submanifold,
thus generalizing a well-known result of S. Kobayashi \cite{kob},
that states that the connected components of the
set of zeros of a Killing vector field on a Riemannian manifold
are totally geodesic submanifolds of even codimension.
A proof of Theorem \ref{umb} was previously given by
D.E. Blair \cite{bla} under the additional restriction (required by
his proof based on the Obata theorem) that the manifold is
compact. Instead, our proof is purely local, as is the proof of the
above mentioned Kobayashi's result.
Very recently, results similar to Theorem \ref{ess} were obtained
independently by M. Lampe in his PhD Thesis \cite{l}.
\section{Proof of the main result}
Let $(M^n,c)$ ($n\ge 3$) be a conformal manifold. We choose a
background Riemannian
metric $g\in c$ and make the usual identifications between
vectors and 1-forms, 2-forms and skew-symmetric endomorphisms etc.,
induced by this metric.
A vector field $\xi$ is called {\em conformal} if the symmetric part of its covariant
derivative with respect to the Levi-Civita connection of $g$
is reduced to its trace:
\begin{equation}\label{1} \nabla_Y\xi=\frac12 Y\,\lrcorner\, d\xi+\varphi Y,\qquad \forall Y\in TM.
\end{equation}
The function $\varphi$ is equal to $-\delta^g\xi/n$, but we will not need this in the sequel.
Let $Z_{\xi}$ and $E_\xi$ denote the zero set and the essential set of $\xi$ respectively.
We have seen that $E_\xi\subset Z_\xi$, so in order to prove Theorem \ref{ess},
it will be enough to show that every limit
point\footnote{A {\em limit point} of a set $S$ in a topological space $X$ is a point $x\in S$ such that every neighborhood of $x$ intersects $S\setminus \{x\}$. The set of limit points of $S$ is denoted by $S'$.} $x$ of $Z_{\xi}$ is not essential,
{\em i.e.} there exists a $\xi$-invariant metric defined locally around $x$.
To do that, we make use of the following criterion by Capocci \cite[Theorem 2.1]{C}, (cf. also
\cite{beig}):
\begin{ath} \cite{C}\label{cap} Let $x\in Z_\xi$ be a zero of the conformal
field $\xi$ on a Riemannian manifold $(M,g)$ of dimension at least $3$ and let $\varphi$ be the function defined by \eqref{1}. Then $\xi$ is a homothetic vector field with
respect to some conformal metric $\tilde g$ defined on a
neighborhood of $x$ if and only if the gradient of
$\varphi$ with respect to $g$ belongs to the image of $\nabla \xi$ at
$x$; moreover, $\xi$ is Killing with respect to $\tilde g$ if, in
addition, $\varphi(x)=0$.
\end{ath}
Theorem \ref{ess} will follow directly from Theorem \ref{cap}, together with:
\begin{ath}\label{fi} Let $x\in Z_{\xi}'$ be a limit point of the
zero set $Z_\xi$ of a conformal vector field $\xi$.
Then the function $\varphi$ defined in \eqref{1} vanishes at $x$ and its gradient
with respect to $g$ belongs to the image of $\nabla \xi$ at $x$.
\end{ath}
\begin{proof} Let $x_k\ne x$ be a sequence of zeros of $\xi$ converging to $x$. In a geodesic chart around $x$, we connect $x$ with
$x_k$ by uniquely defined minimizing geodesics, denoted by
$c_k:[0,t_k]\rightarrow M$, $c_k(0)=x,\ c_k(t_k)=x_k,\ \|\dot{c_k}\|=1$.
The function $f_k:=g(\xi,\dot{c_k})$ admits the following
Taylor-Lagrange expansion:
\begin{equation}\label{elf}f_k(t_k)=f_k(0)+t_kf'_k(\tau_k),\ \mbox{for } \tau_k\in
(0,t_k).\end{equation}
Since $f_k(t_k)=f_k(0)=0$ and $t_k\ne 0$, we have found a sequence of points
$c_k(\tau_k)$, converging to $x$, such that $f'_k(\tau_k)=0$.
On the other hand, equation (\ref{1}) implies that
$f'_k(t)=\varphi(c_k(t))$, therefore $\varphi$ vanishes on a sequence converging
to $x$, thus $\varphi(x)=0$.
For the second part of the theorem, we need to show that $d\varphi_x$ lies in the image of the
skew-symmetric endomorphism $d\xi_x$ of $T_xM$ for every $x\in Z_{\xi}'$.
For a geodesic $c:[0,T]\rightarrow M$ with $c(0)=x$ and $\|\dot{c}\|=1$, we denote by
$\xi'(t)$ and $\xi''(t)$ the derivatives $\nabla_{\dot{c(t)}}\xi$,
respectively $\nabla_{\dot{c(t)}}\nabla_{\dot{c(t)}}\xi$. From \eqref{1} we have
\begin{equation}\label{xip}\xi'(t)=\frac 12
d\xi(\dot{c}(t))+\varphi(c(t))\dot{c}(t),\end{equation}
in particular $\xi'(0)=\frac 12 d\xi(\dot{c}(0))$.
\begin{elem}\label{dxi} For a conformal vector field $\xi$ on a
Riemannian manifold $(M,g)$, such that $\mathcal{L}_\xi g=2\varphi g$, the
following relation holds:
$$\nabla_X d\xi= 2R_{X,\xi}+2d\varphi\wedge X,\qquad \forall X\in TM.$$\end{elem}
\begin{proof} The notations are those from (\ref{1}), which we will use
in the following equivalent form
\begin{equation}\label{dxi2}g(\nabla_A\xi,B)+g(\nabla_B\xi,A)=2\varphi g(A,B),\qquad \forall A,B\in
TM.\end{equation}
Let $x\in M$ and let $X,Y,Z$ be vector fields parallel at $x$. Since
$d\xi(Y,Z)=g(\nabla_Y\xi,Z)-g(\nabla_Z\xi,Y)$, we have at $x$:
\begin{eqnarray*}(\nabla_X
d\xi)(Y,Z)&=&\nabla_X\left(d\xi(Y,Z)\right)=g(\nabla_X\nabla_Y\xi,Z)-g(\nabla_X\nabla_Z\xi,Y)\\
&=&g(R_{X,Y}\xi,Z)-g(R_{X,Z}\xi,Y)+g(\nabla_Y\nabla_X\xi,Z)-g(\nabla_Z\nabla_X\xi,Y).\end{eqnarray*}
We use (\ref{dxi2}) to substitute the last two terms, and the Bianchi
identity for the first two and get
\begin{align*}(\nabla_X d\xi)(Y,Z)\ \ =\ \
\,&g(R_{X,\xi}Y,Z)-g(\nabla_Y\nabla_Z\xi,X)+\nabla_Y(2\varphi
g(X,Z))\\ &+g(\nabla_Z\nabla_Y\xi,X)-\nabla_Z(2\varphi g(Y,X))\\
=\ \ \, &2g(R_{X,\xi}Y,Z)+2d\varphi (Y)g(X,Z)-2d\varphi (Z)g(X,Y).
\end{align*}
\end{proof}
We compute then (for clarity, we omit the argument $t$)
\begin{equation}\label{xisec}\xi''=R_{\dot{c},\xi}\dot{c}+
(d\varphi\wedge\dot{c})(\dot{c})+d\varphi(\dot{c})\dot{c},\end{equation}
in particular, since $\xi_x=0$, we have
\begin{equation}\label{xis0}\xi''(0)=2d\varphi_x(\dot{c}(0))\dot{c}(0)-d\varphi_x.\end{equation}
We come now to the core of our argument. For an arbitrary geodesic $c$
generated by a unit vector in $T_xM$, we estimate the function
$f(t):=g(\xi_{c(t)},\dot{c}(t))$ using the Taylor expansion of order 2:
\begin{equation}\label{ef}f(t)=f(0)+tf'(0)+\frac{t^2}{2}f''(0)+O(t^3).\end{equation}
Here the function $f$ depends on the chosen geodesic, and the error
term $O(t^3)$ is locally bounded (around $x$) by $t^3K$, where $K$ is
a positive constant depending only on the derivatives of $\xi$ but not
on the geodesic $c$. Note
that $\xi_x=0$ implies $f(0)=0$, and
that $f'(0)=\varphi(x)=0$ for any
geodesic $c$.
We choose $c:=c_k$ and $t:=t_k$, and denote by
$f_k(t):=g(\xi_{c_k(t)},\dot{c}(t))$. From \eqref{ef} we infer
\begin{equation} f_k''(0)\longrightarrow 0 \mbox{ for } k\longrightarrow\infty,\end{equation}
Restricting $c_k$ to a subsequence for which $\dot{c}_k(0)$
converges to a unit vector $V\in T_xM$ yields
\begin{equation}\label{dfv}d\varphi_x(V)=0.\end{equation}
\smallskip
In the next step
we estimate $\xi(t):=\xi_{c(t)}$
using a version of the Taylor expansion for vector-valued functions
(here we consider $\xi(.)$ as a function on an interval with values in
$\mathbb{R}^n$, whose components are the components of $\xi$ with respect to
a chosen orthonormal basis parallel along $c$):
\begin{equation}\label{tay}\xi(t)=\xi(0)+t\xi'(0)+\frac{t^2}{2}\xi''(0)+O(t^3).\end{equation}
Note that the function $t\mapsto\xi(t)$ depends on the chosen geodesic $c$,
but the norm of the error term $O(t^3)$ is bounded by $M t^3$ for a
constant $M$
depending only on the derivatives of $\xi$. We now set $c:=c_k$ and
$t=t_k$ and denote the corresponding
function $\xi_{c_k(.)}$ by $\xi_k$. We get
$$0=t_k\xi'_k(0)+\frac{t^2_k}{2}\xi_k''(0)+O(t_k^3),$$
which implies, after taking the quotient by $t_k^2$ and using
(\ref{xip}) and (\ref{xis0}):
\begin{equation}\label{est}\left\|\frac 12 d\xi\left(\frac{\dot{c}_k(0)}{t_k}\right)+
2d\varphi_x(\dot{c}_k(0))\dot{c}_k(0)-d\varphi_x \right\|\le M t_k\ \forall
k\in\mathbb{N}.\end{equation}
Denote now by $V_k$ the vector $\dot{c}_k(0)/2t_k$. The sequence $\{V_k\}$ is
unbounded, but in our relation (\ref{est}) only counts the
projection of $V_k$ on the image of $d\xi_x$, denoted by $W_k:=\pi
V_k$.
Since $\dot{c}_k(0)\longrightarrow V\in T_xM$ and $d\varphi_x(V)=0$ by \eqref{dfv}, we
conclude that the middle term in (\ref{est}) tends to zero, thus
\begin{equation}\label{last}\|d\xi_x(W_k)-d\varphi_x\|\longrightarrow 0,\mbox{ for }
k\longrightarrow\infty.\end{equation}
Since $W_k\perp\ker(d\xi_x)$, it follows that the norm of $W_k$ is
bounded, and we can restrict to a subsequence that converges to
$W\in\mbox{Im}(d\xi_x)$ (Here we consider the kernel and the image
of $d\xi_x$ as the kernel and the image of a skew-symmetric
endomorphism of $T_xM$). It thus follows
$$d\varphi_x=d\xi_x(W),$$
which finishes the proof.
\end{proof}
{\noindent\bf{Remark. }} We have shown that all zeros of $\xi$ which
are not isolated are not essential. Nothing can be said, in general,
about isolated zeros of a conformal vector field, as the following
classical example shows:
Let $\xi$ be the conformal vector field on the round sphere
$S^n\subset \mathbb{R}^{n+1}$, which is
sent through the stereographic projection from one pole $P$ to a
translation vector field on $\mathbb{R}^n$. Then both $\xi$ and its derivative
vanish at $P$, which is the only zero of $\xi$. Therefore, in the
notations of the equation (\ref{1})), $d\xi_P=0$ and $\varphi_P=0$. On the
other hand, $d\varphi_P\ne 0$ (which is implied by the very fact that
$\xi$ is not trivial), so the condition in Theorem \ref{cap} is
violated, which is a proof that $P$ is an essential point for $\xi$.
\section{The zero set of a conformal vector field}
As an application to our main result, we prove
\begin{ath}\label{umb} Let $\xi$ be a conformal vector field on a
Riemannian manifold $(M,g)$ of dimension at least $2$,
and let $Z_{\xi}$ be the zero set of $\xi$ on $M$. Then $Z_{\xi}$ is a
disjoint union of embedded connected totally umbilical submanifolds of $M$, of even codimension when not reduced to a point.
\end{ath}
Let $N\subset M$ be an isometrically embedded submanifold and denote with the same letter $g$ both the metric of $M$ and the induced metric on $N$. Let then $\nabla^M$, $\nabla^N$ be the associated Levi-Civita connections on $M$ and $N$. The $(0,2)$ tensor field $B:T^*N\otimes T^*N\rightarrow \nu(N)$, defined by
$B(X,Y):=\nabla^M_XY-\nabla^N_XY$, where $\nu(N)$ the normal bundle of $N$ in $M$, is called the {\em second fundamental form} of the embedding.
By definition, the submanifold $N$ is called \emph{totally umbilical} if its
second fundamental form is proportional to the metric tensor induced on $N$ in the following sense:
$$B(X,Y)=Hg(X,Y),\ \forall X,Y\in \mathcal{X}(N).$$
The normal vector field $H$ is called the {\em mean curvature} vector field. One-dimensional submanifolds and totally geodesic submanifolds ({\em i.e.} with vanishing second fundamental form) are totally umbilical.
By definition, every
point is considered to be a totally umbilical submanifold of dimension 0.
Using the relation between the Levi-Civita connections $\nabla$, $\nabla'$ of two conformal metrics $g'=e^{2f}g$:
$$\nabla'=\nabla+df\otimes Id +Id\otimes df -g\cdot \mathrm{grad}_gf,$$
one easily sees that if $N$ is totally umbilical with respect to a metric
then
it is totally umbilical with respect to any other conformally equivalent metric. In
particular, $N$ is totally umbilical if there exists a metric $\tilde g$ in the
conformal class, for which $N$ is totally geodesic.
\begin{proof}[Proof of Theorem \ref{umb}]
First, we apply Theorem \ref{ess} to decompose the zero set $Z_{\xi}$ as the
disjoint union of its isolated points $Z^{iso}$ and non-isolated
points $Z':=Z_{\xi}'$, which are inessential, and
conclude that every $x\in Z'$ admits a metric $g_x\in [g]$, defined
locally around $x$, such that $\xi$ is Killing with respect to $g_x$, and
$\xi_x=0$.
The classical result of Kobayashi \cite{kob} implies then that
there is a neighborhood $U_x$ of $x$ such that $Z_{\xi}\cap
U_x=\exp_x(\ker d\xi_x)\cap U_x=\exp_y(\ker d\xi_y)\cap U_x$, for any
$y\in Z_{\xi}\cap U_x$ (here, we consider the dual 1-form
to $\xi$ with respect to $g_x$, and compute its exterior
derivative $d\xi$ at $x$; Also, the exponential map $\exp_x$ is taken
with respect to $g_x$).
In conclusion, every point in $Z_{\xi}$ admits a
neighborhood $U_x$ such that $Z_{\xi}\cap U_x$ is a totally umbilical
submanifold of even codimension. By usual connectedness arguments,
these local totally umbilical submanifolds have a common dimension and
glue together to a global submanifold.
\end{proof}
Note that a classification of totally umbilical submanifolds exists
only for space forms and, more generally, for locally symmetric spaces
(see {\em e.g.} \cite{che}).
The above result can then be a useful tool for producing examples of totally umbilical submanifolds or obstructions to the existence of conformal vector fields on certain Riemannian manifolds.
|
\section{Introduction}
In discussing the operation of gravitational wave (gw)
interferometric detectors \cite{Abramovici, Forward} it is usually
stated that the distance between the beam splitter and the mirrors
at the end of the arms is changed by the gw
\cite{Bluebook,Saulsonbook}. When light beams propagate in the arms,
the change in the proper length of the two arms results in a phase
shift between the light beams that return to the beam splitter. The
phase shift is the physical observable that indicates the presence
of a gw. The above statement is valid if the mirrors are free to
move (along the axis of the arms) and it is expressed in the
laboratory frame of reference, what is referred to as the Local
Lorentz (LL) gauge. It is however much easier to calculate the
resulting phase shift in the Transverse Traceless (TT) gauge. In the
TT gauge the gw has a very simple form and the coordinates of free
particles (i.e. of the mirrors) are not changed in the presence of
the gw, even though their separation does change
\cite{AJP1,AJPSaulson}.
The geometry
of space is defined by the infinitesimal interval \cite{MTW}
\begin{equation}
ds^2 = g_{\mu\nu} dx^{\mu} dx^{\nu}, \qquad \mu, \nu, =
0,1,2,3,
\end{equation}
\noindent where summation over repeated indices is implied. For weak
fields we write the metric tensor as
\begin{equation}
g_{\mu\nu} = \eta_{\mu\nu} + h_{\mu\nu}, \qquad
h_{\mu\nu} \ll 1.
\end{equation}
\noindent The flat-space (Minkowski) metric $\eta_{\mu\nu}$ is
chosen to be
\begin{equation}
\eta_{\mu\nu} = \begin{pmatrix} -1 &0 & 0 & 0\\
0 &1 &0 &0 \\
0 &0 &1 &0 \\
0 &0 &0 &1 \\
\end{pmatrix}.
\end{equation}
Consider the interferometer in the $x-y$ plane and that a
gravitational wave of angular frequency $\Omega$ is incident along
the $z$-axis. In the TT gauge the potential $h_{\mu\nu}$ is given
by the real part of the following expression
\begin{equation}
h_{\mu\nu} = e^{-i(\Omega t + k_{\Omega}z)} \begin{pmatrix} 0 & 0 & 0 & 0\\
0 & h_{+} &h_{\times} &0 \\
0 & h_{\times} &-h_{+} &0 \\
0 & 0 &0 &0 \\
\end{pmatrix}.
\end{equation}
\noindent $h_{+}$ and $h_{\times}$ are the (real) amplitudes of the
``parallel" and ``cross" polarization states of the gw. To be
specific we will place the origin of the coordinates at the beam
splitter and orient the interferometer arms along the $x$- and $y$-
axes, as shown in Fig. (1). In the next section we calculate, in the
TT gauge, the round trip time for the propagation of light from the
beam splitter to the end mirror and back to the origin, in the
presence of a gw. We will then obtain the same result by
considering, again in the TT gauge, the direct interaction of the gw
with the light circulating in the arms.
\begin{figure} [h!]
\centering
\includegraphics[width=70mm,height=70mm]{mich_inter}
\caption{ Configuration of a simple Michelson Interferometer.}
\end{figure}
\section{Conventional calculation}
We will work in the TT gauge so the coordinates of the mirrors are
unchanged by the presence of the gravitational wave. We will
calculate the time it takes for light leaving the beam splitter (the
origin) to reach the end mirror, where it is reflected, and to
return to the the beam origin. For light $ds^2 = 0$, and since the
propagation is in the $x-y$ plane, Eq. (1) reduces to
\begin{equation}
g_{00} c^2 dt^2 + g_{11} dx^2 + g_{22} dy^2 + 2 g_{12} dx dy =
0.\nonumber
\end{equation}
\noindent For propagation only along the $x$-axis, and since we can
set $z = 0$,
\begin{align}
c\left |\frac{dt}{dx}\right | = \left |\frac{g_{11}}{g_{00}}\right
|^{1/2} = (1 + h_{11}(t))^{1/2} \simeq 1 + \frac{1}{2}h_{+}
e^{-i\Omega t} .
\end{align}
\noindent The time interval for a round trip is obtained by
integrating from the origin to the end of the arm, $x = L$, and back
to the origin. If the light leaves at $t = t_0$ and returns at $t =
t_r$
\begin{align}
t_r = t_0 + &\int\limits_0^L \left|\frac{dt}{dx}\right|dx -
\int\limits_L^0 \left|\frac{dt}{dx}\right|dx,
\end{align}
\noindent where the minus sign before the second integral accounts for the
fact that on the return trip $dt/dx = - \left|dt/dx\right|$. In
carrying out the integration we can replace $t$ in the exponential
by $(x/c)$, because any corrections will be of second order in $h$.
We also define the one way travel time in the absence of the gw as
$T = L/c$. Thus
\begin{align}
t_r &= t_0 + \frac{1}{c}\int\limits_0^L
\left[1 + \frac{h_{+}}{2}e^{-i\Omega(t_0 + x/c)}\right]dx
\nonumber \\ & \qquad \ - \frac{1}{c}\int\limits_L^0 \left[1 +
\frac{h_{+}}{2}e^{-i\Omega(t_0 + 2T - x/c)}\right]dx \nonumber \\
&= t_0 + \frac{2L}{c} + \frac{h_{+}}{2i\Omega}
e^{-i\Omega t_0} \left[ 1 - e^{-i2\Omega T}\right]\nonumber\\
&= t_0 + \frac{2L}{c} + h_{+}T e^{-i\Omega (t_0 + T)}\frac{\rm {sin}
\Omega T }{\Omega T}.
\end{align}
\noindent This suggests that we can write
\begin{equation}
h_{11}(t_r) = h_{+}e^{-i\Omega t_r} \simeq h_{+} e^{-i\Omega(t_0 +
2T)}, \nonumber
\end{equation}
\noindent where we have neglected higher order terms in the metric.
As a result, the travel time along the $x$-axis can be expressed as
\begin{equation}
\Delta t_x(t_r) = t_r - t_0 = \frac{2L}{c} + h_{11}(t_r) \frac{L}{c}
\frac{\rm{sin}(\Omega T)}{\Omega T} e^{i\Omega T}.
\end{equation}
\noindent For the $y$-axis we obtain the same result with
$h_{11}(t)$ replaced by $h_{22}(t) = - h_{11}(t)$. Thus the
difference in travel time between the two arms is
\begin{equation}
\Delta t(t) = \Delta t_x(t) - \Delta t_y(t) = 2 [h_{+}e^{-i\Omega
t}] \frac{L}{c} \frac{\rm{sin}(\Omega T)}{\Omega T} e^{i\Omega T}.
\end{equation}\\
A difference in the time of arrival of the light from the two arms,
implies a phase shift between the two fields. If the angular
frequency of the carrier is $\omega_0$, the phase shift
corresponding to a delay $\Delta t$, is $\Delta \phi = \omega_0
\Delta t = c k_0 \Delta t$, where $k_0 = \omega /c$ is the
wavenumber of the carrier. Therefore when the gw is normally
incident on the interferometer plane and when the polarization of
the gw is along the interferometer axes, the observable phase shift
is
\begin{equation}
\Delta \phi(t) = 2 k_0 L h_{+} \frac{\rm{sin}(\Omega T)}{\Omega T}
e^{-i\Omega(t - T)},
\end{equation}
\noindent where following the convention adopted for Eq. (4),
$\Delta \phi (t)$ is given by the real part of Eq. (10). For a more
detailed discussion of this derivation, see \cite{Malik1}.
\section{The field equations in the presence of the gw}
We will now show that the same result can be obtained by considering
the direct interaction of the gravitational wave with the light
propagating in the interferometer arms. In this section we find the
equations for the electromagnetic field in the presence of the gw,
and in the next section we solve the equations for light propagating
in the $x$ and/or $y$ arms. We find the presence of (frequency
modulation) sidebands on the carrier displaced by the gw frequency
$\Omega $. This is equivalent to the phase shift found in Eq. (10).
In flat space Maxwell's inhomogeneous equations can be written in
the manifest covariant form as
\begin{equation}
\partial_{\mu} F^{\mu \nu} = j^{\nu},\label{mweqn}
\end{equation}
where the field strength tensor is defined as
\begin{equation}
F_{\mu\nu} = \partial_{\mu}A_{\nu} - \partial_{\nu}A_{\mu}.\label{fieldstrength}
\end{equation}
In a curved space and in the absence of sources, Eq. \eqref{mweqn} is replaced
by \cite{Landau_Lifshitz}
\begin{equation}
\partial_{\mu}\left(\sqrt{-g}\ g^{\mu \lambda} g^{\nu \rho} F_{\lambda \rho}\right)
= 0,\label{mwcurvedeqn}
\end{equation}
where $g = {\rm{det}}(g_{\mu \nu})$ with the field strength tensor still defined
by Eq. \eqref{fieldstrength}. The dynamical equation \eqref{mwcurvedeqn} can be
derived as the Euler-Lagrange equation from the action described by the Lagrangian density
\begin{align}
{\cal L} & = -\frac{1}{4}\sqrt{-g}\ g^{\mu \lambda} g^{\nu
\rho}F_{\mu \nu} F_{\lambda \rho}
\nonumber\\
& =-\frac{1}{4}\sqrt{-g}\ g^{\mu \lambda} g^{\nu \rho}
\left(\partial_{\mu} A_{\nu} -
\partial_{\nu} A_{\mu}\right)\left(\partial_{\lambda} A_{\rho} - \partial_{\rho}
A_{\lambda}\right).
\end{align}
We will again work in the TT gauge and take the gw to be the same as before.
In the TT gauge, in the weak-field approximation $\sqrt{-g} = 1$, and keeping
only terms linear in $h_{\mu \nu}$,
Eq. \eqref{mwcurvedeqn} reads
\begin{equation}
\partial_{\mu} F^{\mu \nu} - \partial_{\mu}(h^{\mu \lambda} F_{\lambda}^{.\ \nu})
- \partial_{\mu}(h^{\nu \rho} F^{\mu}_{.\ \rho}) = 0.
\end{equation}
\noindent The first term is the same as Eq. \eqref{mweqn} with
$j^{\nu} = 0$, and describes the propagation of the free
electromagnetic field. The next two terms act as sources that give
rise to new electromagnetic fields generated by the interaction of
the gw with the field $F^{\mu \nu}$ of the light circulating in the
arms. For the choice of gw (normal incidence toward the negative
z-axis), expansion of Eq. (14) in $\mathbf{E}\ {\rm{and}}\ \mathbf
{B}$ field components yields the four equations
\begin{align}
\nu = 0:\quad & \mbox{\boldmath$\nabla$} \cdot \mathbf {E} - h_{+}e^{-i(\Omega t + k_{\Omega}z)}(\partial_{x} E_{x} - \partial_{y} E_{y})\nonumber\\
&\quad - h_{\times} e^{-i(\Omega t + k_{\Omega}z)}(\partial_{x} E_{y} + \partial_{y} E_{x}) = 0,
\nonumber \\
\nu = 1:\quad & \partial_{ct} E_{x} - (\mbox{\boldmath$\nabla$} \times \mathbf {B})_{x} + h_{+}e^{-i(\Omega t + k_{\Omega}z)} \nonumber\\
& \quad \times (i\Omega E_{x} + ik_{\Omega}B_{y} - \partial_{ct} E_{x} - \partial_{z} B_{y}) \nonumber \\
& + h_{\times}e^{-i(\Omega t + k_{\Omega}z)} (i\Omega E_{y} + ik_{\Omega}B_x - \partial_{ct} E_{y} + \partial_{z} B_{x})\nonumber\\
&\quad = 0,\nonumber \\
\nu = 2:\quad & \partial_{ct} E_{y} - (\mbox{\boldmath$\nabla$} \times \mathbf {B})_{y} - h_{+}e^{-i(\Omega t + k_{\Omega}z)}\nonumber\\
&\quad \times (i\Omega E_{y} + ik_{\Omega}B_{x} + \partial_{z} B_{x}-\partial_{ct} E_{y})\nonumber \\
& - h_{\times}e^{-i(\Omega t + k_{\Omega}z)} (i\Omega E_{x} + ik_{\Omega}B_y - \partial_{ct} E_{x} - \partial_{z} B_{y})\nonumber\\
&\quad = 0,\nonumber \\
\nu = 3:\quad & \partial_{ct} E_{z} - (\mbox{\boldmath$\nabla$} \times \mathbf {B})_{z} + h_{+}e^{-i(\Omega t + k_{\Omega}z)} (\partial_{x} B_{y} + \partial_{y} B_{x})\nonumber\\
& \quad + h_{\times}e^{-i(\Omega t + k_{\Omega}z)} (\partial_{x}
B_{x} - \partial_{y} B_{y}) = 0.\label{expandedeqn}
\end{align}
\noindent Since $\Omega \ll \omega_{0}$ we can drop the terms in
$\Omega$ and $k_{\Omega}$. We also make use of the condition $z = 0$
and choose \mbox{$h_{+} \not= 0, h_{\times} = 0$.} This leads to the
simplified equations
\begin{align}
\nu = 0:\quad &\mbox{\boldmath$\nabla$} \cdot \mathbf {E} = h_{+}e^{-i\Omega t } (\partial_{x} E_{x} - \partial_{y} E_{y}),\nonumber \\
\nu = 1:\quad &\partial_{ct} E_{x} - (\mbox{\boldmath$\nabla$} \times \mathbf {B})_{x} = h_{+}e^{-i\Omega t } (\partial_{ct} E_{x} + \partial_{z} B_{y}), \nonumber \\
\nu = 2:\quad &\partial_{ct} E_{y} - (\mbox{\boldmath$\nabla$} \times \mathbf {B})_{y} = h_{+}e^{-i\Omega t } (\partial_{z} B_{x}-\partial_{ct} E_{y}), \nonumber \\
\nu = 3:\quad &\partial_{ct} E_{z} - (\mbox{\boldmath$\nabla$}
\times \mathbf {B})_{z} = - h_{+}e^{-i\Omega t } ( \partial_{x}
B_{y} + \partial_{y} B_{x}).\label{lineareqn}
\end{align}
\noindent In the next section we will show how to solve these equations when a plane (light) wave of angular frequency $\omega_{0}$ propagates along the $x$ and/or $y$ axis.
\section{Solution of the equations of motion}
The electromagnetic (em) fields $\mathbf E, \mathbf B$ in the arms consist of the carrier field $\mathbf E_{0},
\mathbf B_{0}$ due to the external source (the laser beam),
and the additional field $\mathbf E_{+}, \mathbf B_{+}$, generated by the source terms in Eq. \eqref{lineareqn},
\begin{equation}
\mathbf E = \mathbf E_{0} + \mathbf E_{+}, \qquad \mathbf B = \mathbf B_{0} + \mathbf B_{+}.
\end{equation}
\noindent We describe the carrier as a plane wave propagating along
the $x$-axis and polarized along the $z$-axis, and of constant amplitude
$A_{0}$
\begin {align}
\mathbf E_{0} & = A_{0}e^{-i(\omega_{0}t - k_{0}x)}\mathbf{ u}_{z}, \nonumber\\
\mathbf B_{0} & = - A_{0}e^{-i(\omega_{0}t - k_{0}x)}\mathbf
{u}_{y}.\label{soln0}
\end{align}
\noindent To the sideband fields $\mathbf E_{+}, \mathbf B_{+}$, we
assign slowly time-dependent amplitudes $E_{+}(t), B_{+}(t)$ and a
phase factor $(\omega_{+}t - k_{0}x)$ where $\omega_{+} = \omega_{0}
+ \Omega$, namely,
\begin {align}
\mathbf E_{+} & = E_{+}(t)e^{-i(\omega_{+}t - k_{0}x)}\mathbf {u}_{z}, \nonumber\\
\mathbf B_{+} & = B_{+}(t)e^{-i(\omega_{+}t - k_{0}x)}\mathbf
{u}_{y}.\label{soln+}
\end{align}
We will solve the equations in \eqref{lineareqn} perturbatively
noting that $E_{+}, B_{+}$ are of order $h$ with respect to $E_{0},
B_{0}$. Thus on the left side of Eq. \eqref{lineareqn} we must use
the full fields $E, B$ but on the right side it suffices to retain
only $E_{0}, B_{0}$. With the choice of Eq. \eqref{soln0} for the
carrier field only the $\nu = 3$ equation is nontrivial; the first
three equations are satisfied automatically with the choices made in
Eqs. (19,20). Writing out the part of order $h$, for the $\nu = 3$
equation in (17), gives
\begin{align}
- \frac{\partial E_{+z}}{c\partial {t}} + \frac{\partial
B_{+y}}{\partial x} & = h_{+}e^{-i\Omega t}\frac{\partial B_{0y}}{\partial x}\nonumber\\
& = -
ih_{+}k_{0}A_{0} e^{-i\omega_{+} t} e^{ik_{0}x}.\label{1st}
\end{align}
\noindent The remaining terms of the equation
\begin{equation}
-\frac{\partial E_{0z}}{c\partial t} + \frac{\partial B_{0y}}{\partial x} = 0,
\end{equation}
\noindent describe the
propagation of the free carrier field. The fields, both carrier and
sideband satisfy Maxwell's dual equation, namely $\partial_{\mu}\tilde F^{\mu \nu} = 0$.
For our choice of direction of propagation and polarization, this condition
reduces to
\begin{equation}
- \frac{\partial E_{+z}}{\partial x} + \frac{\partial B_{+y}}{c\partial t} = 0.\label{2nd}
\end{equation}
\noindent Combining Eqs. \eqref{1st} and \eqref{2nd} leads to wave equations for $E_{+z}, B_{+y}$
\begin{eqnarray}
\frac{\partial ^{2} E_{+z}}{\partial x^{2}} - \frac{\partial ^{2}
E_{+z}}{c^{2}\partial t^{2}} = - h_{+} k_{+} k_{0} A_{0} e^{-i(\omega_{+}t - k_{0} x)},\nonumber\\
\frac{\partial ^{2} B_{+y}}{\partial x^{2}} - \frac{\partial ^{2}
B_{+y}}{c^{2}\partial t^{2}} = + h_{+} k_{0}^{2} A_{0} e^{-i(\omega_{+}t - k_{0} x)},
\end{eqnarray}
\noindent where we have introduced $k_{+} = \omega_{+}/c$.
We now use Eqs. \eqref{soln+} and keep in Eqs. (24) only the first
time derivatives ${dE_{+}(t)}/{dt}, {dB_{+}(t)}/{dt}$ of the
amplitudes, since $E_{+}(t), B_{+}(t)$ vary slowly. We can also
approximate $\omega_{+}^{2}/c^{2} - k_{0}^{2} \simeq 2\Omega
\omega_0/c^{2}$ and find the following two eqs for the slowly
varying amplitudes
\begin{eqnarray}
\frac{dE_{+}}{dt} - i\frac{\omega_0}{\omega_{+}}\Omega E_{+} = i \frac{h_{+}\omega_{0} A_{0}}{2},\nonumber\\
\frac{dB_{+}}{dt} - i\frac{\omega_0}{\omega_{+}}\Omega B_{+} = - i \frac{h_{+}\omega_{0}^{2}A_{0}}
{2\omega_{+}},
\end{eqnarray}
\noindent with solutions
\begin{eqnarray}
E_{+}(t) = \frac{\omega_{+}}{\omega_{0}}\frac{h_{+}\omega_{0} A_{0}}{2 \Omega}
(e^{i\frac{\omega_{0}}{\omega_{+}}\,\Omega t} -1),\nonumber\\
B_{+}(t) = - \frac{h_{+}\omega_{0} A_{0}}{2 \Omega}
(e^{i \frac{\omega_{0}}{\omega_{+}}\,\Omega t} -1).
\end{eqnarray}
\noindent where we assumed the initial conditions $E_{+}(t=0) =
B_{+}(t=0) =0$. Therefore, to the approximation $\Omega \ll \omega_{0}$
that we are using, the sideband fields are
\begin{eqnarray}
\mathbf {E}_{+} = \frac{i\omega_{0}}{\Omega} h_{+} A_{0} e^{\frac{i \Omega
t}{2}} \sin \frac{\Omega t}{2}\, e^{-i(\omega_{+}t -k_{0}x)}\mathbf
{u}_{z},\nonumber\\
\mathbf {B}_{+} = -\frac{i\omega_{0}}{\Omega} h_{+} A_{0} e^{\frac{i \Omega
t}{2}} \sin \frac{\Omega t}{2}\, e^{-i(\omega_{+}t -k_{0}x)}\mathbf
{u}_{y}.
\end{eqnarray}
\noindent The amplitudes $E_{+}(t), B_{+}(t)$ of the sideband fields
are zero at $t=0$ and grow in time as a parametric amplification
process. We will designate them by $A_{+}(t)$. The maximum value of
$A_{+}(t)$ is determined by the losses in the arm cavities, however
the carrier amplitude, $A_{0}$, remains constant. For one round trip
$t=2L/c = 2T$ and the sideband amplitude has the value
\begin{equation}
A_{+}(2T) =i h_{+}L k_{0} A_{0} e^{i\Omega T}\frac{\sin \Omega
T}{\Omega T}.
\end{equation}
We see that after one round trip the amplitude $A_{0}$ of the carrier
acquires a small complex part
\begin{equation}
A_{+}(2T) = iA_{0} \phi_{A+},
\end{equation}
where
\begin{equation}
\phi_{A+} = \phi_{A}e^{i\Omega T} = h_{+} k_{0}L \frac{\sin \Omega T}
{\Omega T} e^{i\Omega T} \ll 1.
\end{equation}
\noindent Thus the amplitude of the propagating light can be written as
\begin{equation}
A = A_{0} + iA_{0}\phi_{A+} \simeq A_{0}e^{i\phi_{A+}},
\end{equation}
\noindent which shows that the phase of the carrier is shifted
by exactly the same amount as was calculated
in section 2. When the carrier propagates in the $y$-arm
the sign of $h$ and thus also of $A_{+}(2T)$ is reversed;
after subtracting the fields returning from the two arms,
we find for their relative phase difference $2\phi_{A+}$, as
we had obtained in Eq. (10).\\
Consider now a gw that depends sinusoidally on time, which we
express at $z = 0$ by
\begin{equation}
h_{+}(t) = h_{+} \cos \Omega t = h_{+}\left(\frac{e^{i\Omega t} +
e^{-i\Omega t}}{2}\right).
\end{equation}
\noindent It is clear that in this case both an upper and lower sideband
will be present. The fields are
\begin{equation}
E_{+} = \frac{iE_{0} \phi_{A+}}{2}\,e^{-i(\omega_{+}t - k_0 x)},
\quad E_{-} = \frac{iE_{0} \phi_{A-}}{2}\,e^{-i(\omega_{-}t - k_0
x)},
\end{equation}
\noindent where $\phi_{A-} = \phi^{*}_{A+} = \phi_{A}e^{-i\Omega
T}$; to see this change \mbox{$\Omega \rightarrow - \Omega$} in
Eq.(30).
At $t=0, x =0$ the two sideband
fields have equal imaginary parts and opposite real parts.
This is the condition that corresponds to phase
(or frequency) modulation when the sidebands are combined with the
carrier
\begin{eqnarray}
E & = & E_{0} e^{-i (\omega_{0} t - k_0 x)}\left(1 +
\frac{i\phi_{A+}}{2}e^{-i\Omega t}
+ \frac{i\phi_{A-}}{2}e^{i\Omega t}\right) \nonumber\\
& = & E_{0} e^{-i (\omega_{0} t - k_0 x)}\left[1 + i\frac{\phi_{A}}{2} (e^{-i\Omega(t - T)}
+ e^{i\Omega(t-T)}) \right]
\nonumber\\
& \simeq & E_{0} e^{-i[ \omega_{0} t - \phi_{A} \cos \Omega (t - T)
- k_0 x]}.
\end{eqnarray}
\noindent In conclusion a sinusoidal gw interacting with a plane wave carrier
(under the appropriate geometry) imposes upper and lower sidebands,
or equivalently contributes a time-dependent phase shift. We
obtained this result by considering the propagation of the carrier
in the space-time of the gw, without referring to the change in the
round-trip time of travel to and from the end mirror.
\section {Discussion}
Our derivation is based on a {\bf{subtle point}}, the use of the TT
gauge. We can use the TT gauge only because the end points of the
light travel (the mirrors) are free. If the mirrors were fixed we
would have to do the calculation in the LL gauge, i.e. in the
laboratory frame of reference. In the LL gauge
\cite{Thorne,Malik2,Pegoraro,Tarabrin} the amplitude of the gw is
modified from the form given by Eq. (4), the relevant part of the
metric being
\begin{equation}
h_{00} = \frac{1}{2c^2} \ddot{h}(t)(x^2 -y^2).
\end{equation}
For low frequency gw's the resulting phase shift when the mirrors
are fixed is much smaller than for free mirrors, and this is why
interferometric gw detectors are constructed with suspended mirrors.
For the details of
the calculation in the LL gauge see ref. \cite{Malik2}. \\
The presence of sidebands has a direct physical interpretation in
terms of absorption and stimulated emission of gravitons from/to the
gw. Since the gw field is highly classical, [the occupation number
for $h \sim 10^{-23}$ and $f \sim 100$ Hz, is $n = N_g/(\lambda
/2\pi)^{3} \sim 10^{33}$], both processes have the same probability.
Absorption leads to the upper sideband, while emission to the lower
one. There is no energy exchange between the optical and
gravitational fields, but
only a phase shift.\\
In our analysis we have assumed that the carrier is a plane wave
propagating toward the positive x-direction, see Eqs. (19). In
practice the carrier is reflected by a mirror at $x=a$, thus the
Electric field should vanish at $x=a$. This condition is satisfied
by writing the fields in the form
\begin {align}
\mathbf E_{0} & = A_{0}[e^{-i(\omega_{0}t - k_{0}x)} - e^{-i(\omega_{0}t
+ k_{0}x - 2 k_{0}a)}]\mathbf{ u}_{z}, \nonumber\\
\mathbf B_{0} & = - A_{0}[e^{-i(\omega_{0}t - k_{0}x)} +
e^{-i(\omega_{0}t + k_{0}x) - 2 k_{0}a}]\mathbf
{u}_{y}.\label{soln00}
\end{align}
If a mirror is also placed at $x=0$ to reflect the carrier and
$k_{0}a =n\pi$, a standing wave is established in the region $0
\leqq x \leqq a$. This does not modify the conclusions that we have
reached. The case when the carrier is a standing wave is treated by
Cooperstock and Faraoni \cite{Cooperstock2} and leads to the same
results as in Eq. (34). The interaction of an electromagnetic and
gravitational field was first discussed by Gertsenshtein and
Pustovoit \cite{Gertsenshtein} in 1963, and
subsequently by others \cite{Caves, Pegoraro, Cooperstock1, Lobo}.\\
We have also shown that a gw will couple to the carrier in a single
arm. The interferometer configuration has been chosen because it is
technically advantageous. The light returning from the two arms is
adjusted to interfere destructively at the detection point in the
absence of a gw. Thus when the gw induces a phase shift, a signal
appears over a null background (excluding noise). By using multiple
traversals in the arms one can increase the effective length, and
thus the phase shift, significantly. The main limitation in
effective arm length (apart from losses in the optics) is related to
the frequency of the gw. We have seen in Eq. (10) that the phase
shift is modulated by the ``form factor" sin$(\Omega T)/\Omega T$,
and therefore $\Omega L/c$ must be kept small.
We thank Dr. R. Weiss for bringing to our attention ref.
\cite{Cooperstock2}. This work was supported in part by
DOE Grant DE-FG02-91ER40685 and
NSF Grant PHY-0456239.
|
\section{Introduction}
In 1957, the issue of the cylindrical gravitational waves of
Einstein and Rosen \cite{Ein} was reviewed by Weber and Wheeler
\cite{We} who were seeking for arguments in favour of the physical
reality of such waves. As a byproduct of their survey, they found
that there could be the circumstances in the theory of general
relativity, under which a space-time is locally flat almost
everywhere but is not Minkowskian in the region of its flatness.
Such a space-time was named conical, and the date of its discovery
is fixed in \cite{We}: in a footnote the authors thank Professor
M~Fierz for permission to quote from his letter of May 14, 1957,
where he showed that conical space-time emerges in the asymptotics
of the imploding-exploding gravitational wave solution at large
distances from the axis of the cylindrical symmetry. A little bit
later the properties of conical space-time were studied in detail by
Marder \cite{Ma1, Ma2} who, in particular, predicted a specific
gravitational lensing effect -- the doubling of the image of objects
located behind the axis of the symmetry (see \cite{Ma2}).
In the same year 1957, Abrikosov \cite{Abr} discovered that a
magnetic vortex can be formed in the type-II superconductors, and
later this result was rederived in a more general context in
relativistic field theory \cite{Nie}. Such string-like structures
denoted as the Abrikosov-Nielsen-Olesen (ANO) vortices arise as
topological defects in the aftermath of phase transitions with
spontaneous breakdown of continuous symmetries; the general
condition of the existence of these structures is that the first
homotopy group of the group space of the broken symmetry group be
nontrivial.
What is the relation between conical spaces and ANO vortices? At
first sight there is nothing, but at a more close look one can
notice that, since the ANO vortex is a topological defect, it is
characterized by nonzero energy distributed along its axis, which in
turn, according to general relativity, is a source of gravity. As
can be shown (for details see the next section), this source makes
the space-time outside the vortex to be conical. Since the squared
Planck length enters as a factor before the stress-energy tensor in
the Einstein-Hilbert equation, the deviations from the Minkowskian
metric are of the order of the squared quotient of the Planck length
to the correlation length, the latter characterizing the size of the
topological defect, i.e. the thickness of the vortex. For
superconductors this quotient is vanishingly small and effects of
the conicity are surely negligible. However topological defects of
the type of ANO vortices may arize in a field which is seemingly
rather different from the condensed matter physics -- in cosmology.
This was realized by Kibble \cite{Ki1,Ki2} and Vilenkin
\cite{Vil1,Vil2} (see also \cite{Zel}), and, from the beginning of
the 1980s, such topological defects in cosmology are known under the
name of cosmic strings. Cosmic strings with the thickness of the
order of the Planck length are definitely ruled out by astrophysical
observations, and there remains a room for cosmic strings with the
thickness which is more than 3.5 orders larger than the Planck
length (see, e.g., \cite{Bat}), although the direct evidence for
their existence is lacking.
In 1959, Aharonov and Bohm \cite{Aha} considered the
quantum-mechanical scattering of a charged particle on a magnetic
vortex and found an effect that does not depend on the depth of
penetration of the charged particle into the region of the vortex
flux. Thus, it was demonstrated for the first time that the
quantum-mechanical motion of charged particles can be affected by
the magnetic flux from the impenetrable for the particles region.
This effect which is alien to classical physics has a great impact
on the development of various fields in quantum physics, ranging
from particle physics and cosmology to condensed matter and
mesoscopic physics (see, e.g., reviews \cite{Ola,Pes,Kri}).
In the late 1980s, the quantum-mechanical scattering of a test
particle in conical space was considered by t'Hooft \cite{Ho} and
Jackiw et al \cite{Des, Sou}; later the consideration was extended
to the case of a magnetic vortex placed along the axis of conical
space, i.e. to the quantum-mechanical scattering on a cosmic string
\cite{Si2}. It should be noted that in the above works, as well as
in \cite{Aha}, the effects of the thickness of the strings were
neglected. This shortcoming was remedied, and the finite-thickness
effects were taken into account both for the vortex in ordinary flat
space \cite{Ola} and for the vortex in conical space \cite{Si5} (see
also \cite{SiV}). Paradoxical peculiarities of the Aharonov-Bohm
effect in conical space, including the issue of the
quantum-classical correspondence, are discussed in the present
paper.
\section{Abrikosov-Nielsen-Olesen vortex and conical space}
Let us start with the lagrangian of the Abelian Higgs model
\begin{equation}
\fl L=-\frac14F_{\rho \rho'}F^{\rho \rho'}-\left[\left(\partial^\rho-{\rm i}e_{\rm H}A^\rho\right)\psi_{\rm H}\right]^*
\left[\left(\partial_\rho-{\rm i}e_{\rm H}A_\rho\right)\psi_{\rm H}\right]-\frac \lambda 4\left(\psi_{\rm H}^*\psi_{\rm H}-
\frac{\sigma^2}{2}\right)^2,\label{eq1}
\end{equation}
where units $c=\hbar=1$ are used and the metric signature is chosen
as $(-1,\,1,\,1,\,1)$. The ground state of the model is
characterized by the nonzero vacuum expectation value of the
absolute value of the complex scalar Higgs field
\begin{equation}
\langle{\rm vac}||\psi_{\rm H}||{\rm vac}\rangle=\frac{1}{\sqrt{2}}\sigma,\label{eq2}
\end{equation}
where the value of $\sigma$ is implied to be real positive. Writing
complex field $\psi_{\rm H}$ in terms of two real fields $\chi_{\rm
H}$ and $\widetilde{\chi}_{\rm H}$
\begin{equation}
\psi_{\rm H}(x)=\frac{1}{\sqrt{2}}[\sigma+\chi_{\rm H}(x)]e^{{\rm i}\widetilde{\chi}_{\rm H}(x)},\label{eq3}
\end{equation}
and eliminating field $\widetilde{\chi}_H$ by gauge transformation,
$A_\rho\rightarrow A_\rho+e^{-1}_{\rm H}
\partial_\rho\widetilde{\chi}_{\rm H}$ , one can present $L$ (1) in
the form
\begin{equation}
L=-\frac14F_{\rho \rho'}F^{\rho \rho'}-\frac 12m_{\rm A}^2A^\rho A_\rho -\frac 12(\partial^\rho\chi_{\rm H})
(\partial_\rho\chi_{\rm H})-\frac 12m_{\rm H}^2\chi_{\rm H}^2+\ldots,\label{eq4}
\end{equation}
where dots correspond to cubic and quartic self-interaction terms of
$\chi_H$ and to a quartic interaction term of $\chi_{\rm H}$ with
$A_\rho$, and
\begin{equation}
m_{\rm A}^2=e_{\rm H}^2\sigma^2,\qquad m^2_{\rm H}=\frac 12\lambda\sigma^2.\label{eq5}
\end{equation}
Thus, the physical content of the model is the vector particle with
mass $m_{\rm A}$ and the scalar particle with mass $m_{\rm H}$.
A static cylindrically symmetric solution in the model is given by
configuration \cite{Nie}
\begin{equation}
\psi_{\rm H}=\frac {1}{\sqrt{2}}\sigma\tau_{\rm H}(r)e^{{\rm i}n\varphi},\label{eq6}
\end{equation}
\begin{equation}
A_\varphi=\frac{n}{e_{\rm H}}[\tau_{\rm A}(r)]^2,\qquad A_0=A_r=A_3=0,\label{eq7}
\end{equation}
where cylindrical $(r,\,\varphi,\,x^3)$ coordinates are used, $n\in
\mathbb{Z}$ ($\mathbb{Z}$ is the set of integer numbers), and
$\tau_H$ and $\tau_{\rm A}$ satisfy the system of nonlinear
differential equations
\begin{equation}
\left\{\begin{array}{l}
r^{-1}\partial_rr\partial_r\tau_{\rm H}+\frac 12 m_{\rm H}^2(1-\tau_{\rm H}^2)\tau_{\rm H}-n^2r^{-2}(1-\tau_{\rm A}^2)^2\tau_{\rm H}=0,\\
r\partial_r r^{-1}\partial_r\tau_{\rm A}^2+m_{\rm A}^2\tau_{\rm H}^2(1-\tau_{\rm A}^2)=0,
\end{array}\right.
\label{eq8}
\end{equation}
with boundary conditions
\begin{equation}
\tau_{\rm H}(0)=\tau_{\rm A}(0)=0,\qquad \tau_{\rm H}(\infty)=\tau_{\rm A}(\infty)=1.\label{eq9}
\end{equation}
The analytical form for the solution to (8) and (9) is unknown, but
the extensive analysis involving the use of rigorous methods (see
\cite{Jaf}) and numerical calculations yields that $\tau_{\rm H}$
and $\tau_{\rm A}$ tend to zero at the origin as $\tau_{\rm
H}=a_{\rm H}r^{|n|}$ and $\tau_{\rm A}=a_{\rm A}r$, while
approaching exponentially fast to the unity value as
$\tau_H=1-b_{\rm H}r^{-1/2}\exp(-m_{\rm H}r)$ and $\tau_{\rm
A}=1-b_{\rm A}r^{1/2}\exp(-m_{\rm A}r)$. The only nonvanishing
component of gauge field tensor $F_{\rho\rho'}$ is
\begin{equation}
B^3\equiv r^{-1}F_{r\varphi}=r^{-1}\partial_rA_\varphi=2n(e_{\rm H} r)^{-1}\tau_{\rm A}(\partial_r\tau_{\rm A}).\label{eq10}
\end{equation}
Thus, one notes that at large distances from the symmetry axis (at
$r>m_{\rm H}^{-1}$ and $r>m_{\rm A}^{-1}$) the solution to (8) and
(9) corresponds to the ground state: $\psi_{\rm
H}=\frac{1}{\sqrt{2}}\sigma e^{{\rm i}n\varphi}$ and $B^3=0$. The
solution is characterized by two cores: the one (where the Higgs
field differs from its vacuum value) has the transverse size of the
order of correlation length $r_{\rm H}=m_{\rm H}^{-1}$, and the
other one (where the gauge field strength is nonzero) has the
transverse size of the order of penetration depth $r_{\rm A}=m_{\rm
A}^{-1}$. The value of quotient $\kappa\equiv r_{\rm A}/r_{\rm
H}=e_{\rm H}^{-1}\sqrt{\lambda/2}$ which is known as the
Ginzburg-Landau parameter distinguishes between the type-I
($\kappa<1/\sqrt{2}$) and the type-II ($\kappa>1/\sqrt{2}$)
superconductors. The solution to (8) and (9) in the case of the
type-II superconductors is known as the Abrikosov vortex, while the
solution in general case of either $\kappa>1/\sqrt{2}$ or
$\kappa<1/\sqrt{2}$ may be denoted as the Abrikosov-Nielsen-Olesen
(ANO) vortex. The ground state manifold, i.e. the spatial region
outside the vortex, is not simply connected; the first homotopy
group $\pi_1$ is nontrivial, $\pi_1=\mathbb{Z}$, and thus the
vortices are characterized by winding number $n\in \mathbb{Z}$, see
(6) and (7).
The stress-energy tensor corresponding to the ANO vortex has
diagonal nonvanishing components only, and
\begin{eqnarray}
\fl T_{00}=-T_{33}=\frac 12(B^3)^2+r^{-2}|(\partial_\varphi-ie_{\rm H} A_\varphi)\psi_{\rm H}|^2+
|\partial_r\psi_{\rm H}|^2
+\frac{\lambda}{4}(|\psi_{\rm H}|^2-\frac{\sigma^2}{2})^2=\nonumber \\ \fl =\frac 12\sigma^2
\left\{(\partial_r\tau_{\rm H})^2+\frac 14 m_{\rm H}^2(1-\tau_{\rm H}^2)^2+n^2r^{-2}
[4m_{\rm A}^{-2}\tau_{\rm A}^2(\partial_r\tau_{\rm A})^2+\tau_{\rm H}^2(1-\tau_{\rm A}^2)^2]\right\}. \label{eq11}
\end{eqnarray}
As to $T_{rr}$ and $r^{-2}$ $T_{\varphi \varphi}$, they are negative
with their absolute values being much smaller than $T_{00}$ (see,
e.g., \cite{Gar}).
Using (8), relation (11) is recast into the form
\begin{equation}
\fl T_{00}=-T_{33}=\frac 12\sigma^2[r^{-1}\partial_r(r\tau_{\rm H}\partial_r\tau_{\rm H})+
\frac 14 m_{\rm H}^2(1-\tau_{\rm H}^4)+4n^2m_{\rm A}^{-2}r^{-2}\tau_{\rm A}^2(\partial_r\tau_{\rm A})^2].\label{eq12}
\end{equation}
In the square brackets on the right-hand side of (12), the first two
terms are nonvanishing in the core of the order of correlation
length $r_{\rm H}$ and the third term is nonvanishing in the core of
the order of penetration depth $r_{\rm A}$.
The stress-energy tensor is a source of gravity according to the
Einstein--Hilbert equation
\begin{equation}
R_{\rho\rho'}-\frac 12 g_{\rho\rho'}R=8\pi GT_{\rho\rho'},\label{eq13}
\end{equation}
where $R_{\rho\rho'}$ is the Ricci tensor,
$R=g^{\rho\rho'}R_{\rho\rho'}$ is the scalar curvature, $G=l^2_{\rm
Pl}$ is the gravitational constant ($l_{\rm Pl}$ is the Planck
length); we use the notations adopted in \cite{Mis}. Taking the
trace over Lorentz indices in (13), one gets that the space-time
region of the vortex core is characterized by the positive scalar
curvature, $R>16\pi GT_{00}$, since $T_{00}$ (12) is positive there.
Space-time outside the vortex core is flat ($R=0$) but
non-Minkowskian. To see this, let us first integrate (12) over the
transverse spatial dimensions and get the energy per unit length of
the vortex
\begin{equation}
\mu=\int\limits_{0}^{2\pi}{\rm d}\varphi\int\limits_{0}^{\infty}{\rm d}r\,rT_{00}=\pi\sigma^2(I_{\rm H}+I_{\rm A}),\label{eq14}
\end{equation}
where
\begin{equation}
I_{\rm H}=\frac 14\int\limits_{0}^{\infty}{\rm d}u\,u(1-\bar{\tau}_{\rm H}^4),\qquad I_{\rm A}=4n^2\int\limits_{0}^{\infty}
\frac{{\rm d}u}{u}\bar{\tau}_{\rm A}^2(\partial_u\bar{\tau}_{\rm A})^2,\label{eq15}
\end{equation}
and the functions of dimensionless variables are introduced:
$\bar{\tau}_{\rm H}(m_{\rm H}r)\equiv\tau_{\rm H}(r)$ and
$\bar{\tau}_{\rm A}(m_{\rm A}r)\equiv \tau_{\rm A}(r)$. The
integrands in integrals (15) are damped as $e^{-u}$ at large values
of $u$, and, therefore, the upper limit of integration in (15) is of
order of unity rather than infinity. One can conclude that the
dependence of linear energy density $\mu$ (14) on parameter $\kappa$
is rather weak. In order to solve (13) outside the vortex core, it
suffices to use the approximation neglecting the transverse size of
the core. Then the stress-energy tensor is expressed in terms of
$\mu$ as
\begin{equation}
T_{00}=-T_{33}=\mu\frac{\delta(r)}{r}\Delta(\varphi),\qquad T_{rr}=T_{\varphi\varphi}=0,\label{eq16}
\end{equation}
where
$\Delta(\varphi)=(2\pi)^{-1}\sum\limits_{n\in\mathbb{Z}}e^{{\rm
i}n\varphi}$ is the delta-function for the compact (angular)
variable. Solving (13) with $T_{\rho\rho'}$ in the form of (16), one
gets the metric outside the vortex core, which is given by squared
length element
\begin{equation}
\fl {\rm d}s^2=-{\rm d}t^2+(1-4G\mu)^{-1}{\rm d}r^2+(1-4G\mu)r^2{\rm d}
\varphi^2+({\rm d}x^3)^2=-{\rm d}t^2+{\rm d}\tilde{r}^2+\tilde{r}^2{\rm d}\tilde{\varphi}^2
+({\rm d}x^3)^2,\label{eq17}
\end{equation}
where
$$
\tilde{r}=r(1-4G\mu)^{-1/2},\qquad 0<\tilde{\varphi}<2\pi(1-4G\mu).
$$
This is the metric of conical space: a surface which is transverse
to the axis of the vortex is isometric to the surface of a cone with
the deficit angle equal to $8\pi G\mu$.
In view of (14), the value of the linear energy density can be
estimated as $\mu\approx \pi\sigma^2$. The Abelian Higgs model (1)
contains two dimensionless parameters, $\lambda$ and $e_{\rm H}$,
and, by varying their values, one gets the variety of
superconductors of types I and II. One can fix the value of one
parameter, say $\lambda$, and then the variety of superconductors is
obtained by varying the value of other parameter, $e_{\rm H}$. By
fixing $\lambda =2\pi$, one gets $\mu\approx m_{\rm H}^2$, and then
the deviation from the Minkowskian metric is estimated as
$G\mu\approx(l_{\rm Pl}/r_{\rm H})^2$, as it has been announced in
Introduction.
A more consistent treatment involves the analysis of the full system
of coupled equations for the metric and the vortex-forming gauge and
Higgs fields \cite{Gar}; it yields the same results as presented
above.
To conclude this section, we list the global (spatial-point
independent) parameters of the ANO vortex: flux of the gauge field
strength,
\begin{equation}
\Phi=\int\limits_{\rm core}{\rm d}\sigma\,B^3,\label{eq18}
\end{equation}
and linear energy density (compare with (14) and (15)),
\begin{equation}
\mu=\int\limits_{\rm core}{\rm d}\sigma\,T_{00},\label{eq19}
\end{equation}
where the integration is over the transverse section of the core of
the vortex. The flux is directly related to the gauge-Higgs
coupling, $\Phi=2\pi\hbar {\rm c}ne_{\rm H}^{-1}$ (see (10)), while
the density can be estimated as $\mu\approx m^2_{\rm H}{\rm
c}^3/\hbar$ or $\mu\approx \hbar {\rm c}/r_{\rm H}^2$, where
constants ${\rm c}$ and $\hbar$ are recovered. One more parameter
should be introduced, and this is the transverse radius of the
vortex core,
\begin{equation}
r_{\rm c}={\rm max}\{r_{\rm H},\,r_{\rm A}\},\label{eq20}
\end{equation}
i.e. $r_{\rm H}$ for the type-I superconductors and $r_{\rm A}$ for
the type-II superconductors.
\section{Quantum-mechanical scattering in conical space}
We shall study the quantum-mechanical scattering of a
nonrelativistic test particle by an ANO vortex. Since the motion of
the particle along the vortex axis is free, we need only to consider
the two-dimensional motion on the surface which is orthogonal to the
vortex axis. The Schr\"{o}dinger equation for the wave function
describing the stationary scattering state has the form
\begin{equation}
{\rm H}\psi(r,\,\varphi)=\frac{\hbar^2k^2}{2m}\psi(r,\,\varphi),\label{eq21}
\end{equation}
where $m$ is the particle mass and $k$ is the absolute value of the
particle wave vector.
The conical nature of space outside the vortex core is characterized
by dimensionless parameter
\begin{equation}
\eta=4G\mu{\rm c}^{-4},\label{eq22}
\end{equation}
where, as well as in (21), constants ${\rm c}$ and $\hbar$ are
recovered. As it has been already mentioned, this parameter is
vanishingly small for vortices in superconductors, while for cosmic
strings it is restricted to the range $0<\eta<4\cdot 10^{-7}$
\cite{Bat}. However, it may appear that a more wide range of $\eta$
is of physical interest: in particular, the topological defects in
graphene correspond to carbon monolayer nanocones with deficit
angles $2\pi \eta$ being positive and negative integer multiples of
$\pi/3$ \cite{Si7,Si8}. In view of this, we make our consideration
as general as possible by extending the range of values of $\eta$ to
$1>\eta>-\infty$\footnote{If one considers a set of noncompact
simply connected surfaces imbedded in three-dimensional Euclidean
space, then $2\geq \eta>-\infty$ and asymptotically conical surfaces
($1>\eta>-\infty$), as well as an asymptotically cylindrical surface
($\eta=1$), have infinite area, whereas surfaces with $2\geq\eta >1$
have finite area (e.g., a surface with $\eta=2$ is a sphere with a
puncture). A cylindrically symmetric space, where a surface
orthogonal to the symmetry axis is asymptotically conical, can be
denoted, in general, as a conical space. Thus such spaces are
characterized by $\eta$ from the range $1>\eta>-\infty$.}.
The behaviour of the Schr\"{o}dinger hamiltonian at large distances
from the scattering centre is crucial for the construction of
scattering theory, therefore it suffices at first to write down the
hamiltonian outside the vortex core
\begin{equation}
{\rm H}=-\frac{\hbar^2}{2m}\left[\partial_r^2+\frac 1r\partial_r+\frac{1}{(1-\eta)^2r^2}
\left(\partial_\varphi-{\rm i\frac{\Phi}{\Phi_0}}\right)^2\right], \quad r>r_{\rm c},\label{eq23}
\end{equation}
where $\Phi_0=2\pi\hbar{\rm c}{\rm e}^{-1}$ is the London flux
quantum. The quantum-mechanical particle is coupled to the
vortex-forming gauge field with constant ${\rm e}$ which, in
general, differs from ${\rm e}_{\rm H}$; the case of ${\rm e}={\rm
e}_{\rm H}/2$ corresponds to the Bardeen-Cooper-Schrieffer model of
superconductivity, when the condensate of Cooper pairs is described
phenomenologically by the Higgs field. A vortex in the type-II
superconductors has the flux equal to one semifluxon, $\Phi=\frac
12\Phi_0$, while a vortex in the type-I superconductors can have the
flux equal to an integer multiple of a semifluxon, $\Phi=\frac
n2\Phi_0$ ($n\in\mathbb{Z}$).
Hamiltonian (23) should be compared with the hamiltonian in the
absence of the vortex (i.e. at $\Phi=0$ and $\eta=0$):
\begin{equation}
{\rm H}_0=-\frac{\hbar^2}{2m}\left(\partial_r^2+\frac 1r\partial_r+\frac{1}{r^2}\partial_\varphi^2\right).\label{eq24}
\end{equation}
Thus the interaction is given by difference between (23) and (24),
which can be written in the form
\begin{equation}
{\rm H}-{\rm H}_0=v({\bf x})+v^j({\bf x})(-{\rm i}\frac{\partial}{\partial x^j})+v^{jj'}({\bf x})
\left(-\frac{\partial^2}{\partial x^j\partial x^{j'}}\right),\label{eq25}
\end{equation}
where we have introduced notations ${\bf x}=(x^1,\,x^2)$, $x^1=r\cos
\varphi$, $x^2=r\sin \varphi$, and $j,\,j'=1,\,2$.
It should be noted that interaction of the form of (25) is of short
range, if coefficient functions $v$, $v^j$, and $v^{jj'}$ decrease
as $O(r^{-1-\varepsilon})$ at $r\rightarrow\infty$
($\varepsilon>0$), and then scattering theory can be constructed in
the usual way (see, e.g., \cite{Ree}). However, even for particle
scattering by a purely magnetic vortex ($\Phi\neq 0$ and $\eta=0$)
interaction (25) is of long range since coefficient function $v^j$
decreases as $O(r^{-1})$ at $r\rightarrow \infty$. Because of the
long-range nature of the interaction in this case, it is impossible
to choose a plane wave as the incident wave, as it has been noted by
Aharonov and Bohm \cite{Aha}. Nevertheless, it is possible to
construct scattering theory in this case and obtain in its framework
the Aharonov-Bohm scattering amplitude (see \cite{Rui}).
H\"{o}rmander \cite{Hor} studied a class of interactions of the form
(25) containing both a short-range part and a long-range part
characterized by real coefficient functions that decrease in the
limit $r\rightarrow\infty$ as $O(r^{-\varepsilon})$
($0<\varepsilon\leq 1$). He formulated certain additional
requirements under which scattering theory can be constructed, and,
as he notes in his monograph \cite{Hor}, "the existence of modified
wave operators is proved under the weakest sufficient conditions
among all those known at the present time".
H\"{o}rmander's conditions are satisfied by the interaction in the
problem of scattering by a purely magnetic vortex,
$$
v\sim O(r^{-2}) \quad {\rm and}\quad v^j\sim O(r^{-1}),\quad r\rightarrow\infty
$$
($v$ and $v^j$ are real, and $v^{jj'}=0$), and, for instance, by the
interaction in the problem of scattering by a Coulomb centre,
$$
v\sim O(r^{-1}),\quad r\rightarrow\infty
$$
($v$ is real, and $v^j=v^{jj'}=0$). On the contrary, the interaction
in the problem of scattering by a vortex in conical space ($\Phi\neq
0$, $\eta\neq 0$) does not satisfy H\"{o}rmander's conditions:
\begin{equation}
v\sim O(r^{-2}),\quad v^j\sim O(r^{-1}) \quad {\rm and} \quad v^{jj'}\sim O(1), \quad r\rightarrow\infty,\label{eq26}
\end{equation}
where $v^j$, in contrast to $v$ and $v^{jj'}$, is a complex function
(more precisely, the imaginary part of $v^j$ of order $r^{-1}$ is
due to the nondecrease of real quantity $v^{jj'}$ in the limit
$r\rightarrow\infty$). Nevertheless, even in this last case
scattering theory can be constructed, and this has been done in
\cite{Si5}, based on earlier works \cite{Des,Sou,Si2}.
According to this theory, the scattering matrix in the wave vector
representation is
\begin{eqnarray}
\fl S(k,\,\varphi;\,k',\,\varphi ')=\frac 12\frac{\delta(k-k')}{\sqrt{kk'}}{\rm e}^{2{\rm i}k(r_{\rm c}-\xi_{\rm c})}
\left\{\Delta\left(\varphi-\varphi '+\frac{\eta\pi}{1-\eta}\right)\exp\left[-\frac{{\rm i}\Phi\pi}{\Phi_0(1-\eta)}
\right]+\right. \nonumber \\
\fl \left.+\Delta\left(\varphi-\varphi '-\frac{\eta\pi}{1-\eta}\right)\exp\left[\frac{{\rm i}\Phi\pi}{\Phi_0(1-\eta)}
\right]\right\}+\delta(k-k')\frac{e^{{\rm i}\pi/4}}{\sqrt{2\pi k}}f(k,\,\varphi-\varphi '),\label{eq27}
\end{eqnarray}
where the final (${\bf k}$) and initial (${\bf k}'$) two-dimensional
wave vectors of the particle are written in polar variables, and
$\xi_{\rm c}=\int\limits_{0}^{r_{\rm c}}ds$ is the geodesic radius
of the vortex core. All the delta-functions of angular variables are
enclosed in the figure brackets, whereas the transition matrix (the
last term in (27)) is free of such delta-functions. Note that in the
case of the short-range interaction one has $2\Delta(\varphi-\varphi
')$ instead of the figure brackets in (27). Thus, one can see that,
due to the long-range nature of the interaction, even the
conventional relation between the scattering matrix and the
transition matrix is changed, involving now a distorted unity matrix
(first term in (27)) instead of the usual one,
$\delta(k-k')\Delta(\varphi-\varphi ')(kk')^{-1/2}$.
The transition matrix contains the scattering amplitude ($f$) which
is given by expression
\begin{equation}
f(k,\,\varphi-\varphi ')={\rm e}^{2{\rm i}k(r_{\rm c}-\xi_{\rm c})}f_0(k,\,\varphi-\varphi ')
+f_{\rm c}(k,\,\varphi-\varphi '),\label{eq28}
\end{equation}
where
\begin{equation}
f_0(k,\,\varphi)=-\frac{{\rm e}^{{\rm i}\pi/4}}{\sqrt{2\pi k}}\sum\limits_{n\in\mathbb{Z}}\exp[{\rm i}n
(\varphi-\pi)]\sin(\alpha_n\pi),\label{eq29}
\end{equation}
\begin{eqnarray}
f_{\rm c}(k,\,\varphi)=-\exp[2{\rm i}k(r_{\rm c}-\xi_{\rm c})-{\rm i}\pi/4]\sqrt{\frac{2}{\pi k}}\times \nonumber \\
\times\sum\limits_{n\in\mathbb{Z}}\exp[{\rm i}n(\varphi-\pi)-{\rm i}\alpha_n\pi]
\frac{W[\sqrt{\xi_{\rm c}}\kappa_n(\xi_{\rm c},\,k),\,\, \sqrt{r_{\rm c}}J_{\alpha_n}(kr_{\rm c})]}
{W[\sqrt{\xi_{\rm c}}\kappa_n(\xi_{\rm c},\,k),\,\, \sqrt{r_{\rm c}}H_{\alpha_n}^{(1)}(kr_{\rm c})]},\label{eq30}
\end{eqnarray}
\begin{equation}
\alpha_n=|n-\Phi/\Phi_0|(1-\eta)^{-1},\label{eq31}
\end{equation}
$J_\nu(u)$ and $H_\nu^{(1)}(u)$ are the Bessel and the first-kind
Hankel functions of order $\nu$,
$\kappa_n\left(\int\limits_{0}^{r}ds,\,k\right)$ is the partial wave
solution which is unique and regular inside the vortex core, and the
Wronskian of functions ${\cal F}^{(1)}(\xi_{\rm c})$ and ${\cal
F}^{(2)}(r_{\rm c})$ is defined as
$$
W\left[{\cal F}^{(1)}(\xi_{\rm c}),\,{\cal F}^{(2)}(r_{\rm c})\right]={\cal F}^{(1)}(\xi_{\rm c})
\left[\partial_r{\cal F}^{(2)}(r)\right]|_{r=r_{\rm c}}-
\left[\partial_r{\cal F}^{(1)}(r)\right]|_{r=\xi_{\rm c}}{\cal F}^{(2)}(r_{\rm c}).
$$
It is also instructive to present the $r\rightarrow\infty$
asymptotics of the scattering wave solution to the Schr\"{o}dinger
equation (21)
\begin{eqnarray}
\fl \psi({\bf x},\,{\bf k}')=(2\pi)^{-1}\sum\limits_{l}\exp\{-{\rm i}kr\cos[(1-\eta)(\varphi-\varphi '
-\pi+2l\pi)]\}\times \nonumber \\
\fl \times\exp[{\rm i}\Phi\Phi_0^{-1}(\varphi-\varphi '-\pi+2l\pi)]+f(k,\,\varphi-\varphi ')
[2\pi(1-\eta)\sqrt{r}]^{-1}\exp({\rm i}kr)+O(r^{-1}),\label{eq32}
\end{eqnarray}
where ${\bf x}=(r\,\cos\varphi,\,r\sin\varphi)$, ${\bf
k}'=(k\,\cos\varphi ',\,k\,\sin\varphi ')$, and the summation is
over integer values of $l$ that satisfy condition
\begin{equation}
\frac{\varphi '-\varphi}{2\pi}-\frac 12\frac{\eta}{1-\eta}<l<\frac{\varphi '-\varphi}{2\pi}+
1+\frac 12\frac{\eta}{1-\eta}.\label{eq33}
\end{equation}
Scattering amplitude $f(k,\,\varphi-\varphi ')$ enters (32) as the
factor before outgoing wave $[2\pi(1-\eta)\sqrt{r}]^{-1}\exp({\rm
i}kr)$. In the case of the short-range interaction, the term of
order $O(1)$ is plane wave $(2\pi)^{-1}\exp[{\rm
i}kr\cos(\varphi-\varphi ')]$ which is interpreted as the incident
wave. In the case of scattering by a purely magnetic vortex, the
incident wave is distorted, differing from the plane wave and taking
the form $(2\pi)^{-1}\exp[{\rm i}kr\cos(\varphi-\varphi ')]\exp[{\rm
i}\Phi\Phi_0^{-1}(\varphi-\varphi '-\pi)]$ \cite{Aha}. When space is
conical, the distortion is much stronger and the incident wave is
given by the finite sum which is of order $O(1)$ in (32); the
distortion of the incident wave in conical space was first obtained
in \cite{Ho,Des}.
The independent of $r_{\rm c}$ part of the scattering amplitude,
$f_0(k,\,\varphi)$ (29), is determined by the sum which can be
exactly taken \cite{Si2} yielding:
\begin{eqnarray}
\fl f_0(k,\varphi)\!=\!-\frac{{\rm e}^{{\rm i}\pi/4}}{2\sqrt{2\pi k}}\Biggl\{\!\exp
\left[{\rm i}\left[\!\!\left[\frac{\Phi}{\Phi_0}\right]\!\!\right]\!\left(\varphi\!+\!
\frac{\eta\pi}{1\!-\!\eta}\right)\!-\frac{{\rm i}\Phi\pi}{\Phi_0(1\!-\!\eta)}\right]\left[\cot
\left(\frac 12\left(\varphi+\frac{\eta\pi}{1\!-\!\eta}\right)\right)\!+\!{\rm i}\right]\Biggr.\!- \nonumber \\
\fl \Biggl.-\exp\left[{\rm i}\left[\!\!\left[\frac{\Phi}{\Phi_0}\right]\!\!\right]\left(\varphi-
\frac{\eta\pi}{1-\eta}\right)+\frac{{\rm i}\Phi\pi}{\Phi_0(1-\eta)}\right]\left[\cot
\left(\frac 12\left(\varphi-\frac{\eta\pi}{1-\eta}\right)\right)+{\rm i}\right]\Biggr\},\label{eq34}
\end{eqnarray}
where $[\![u]\!]$ denotes the integer part of quantity $u$ (i.e. the
integer which is less or equal to $u$). In the limit $r_{\rm
c}\rightarrow 0$ one gets $\xi_{\rm c}\rightarrow 0$ and $f_{\rm
c}(k,\,\varphi)\rightarrow 0$, since function $\kappa_n$ is regular,
while $J_{\alpha_n}$ is vanishing and $H_{\alpha_n}^{(1)}$ is
divergent at the origin (to be more specific, $f_{\rm
c}(k,\,\varphi)$ decreases at least as $O[\ln^{-1}(kr_{\rm c})]$ at
$r_{\rm c}\rightarrow 0$ \cite{Si5}). Thus $f_0(k,\,\varphi)$ (34)
is the amplitude of scattering by an idealized (singular) vortex of
zero thickness. The long-range nature of interaction exhibits itself
in the divergence of scattering amplitude (34) in two directions
which are symmetric with respect to the forward direction,
$\varphi=\pm \eta\pi(1-\eta)^{-1}$; note that the amplitude of
scattering by a purely magnetic vortex ($\eta=0$) diverges in the
forward direction, $\varphi=0$ \cite{Aha}.
Since dimensionless quantity $\sqrt{k}f_{\rm c}(k,\,\varphi)$
depends on $r_{\rm c}$ through dimensionless product $kr_{\rm c}$,
the limit $r_{\rm c}\rightarrow 0$ is the same as the limit
$k\rightarrow 0$. Therefore, in the long-wavelength limit
($k\rightarrow 0$), $f_{\rm c}(k,\,\varphi)$ is negligible as
compared to $f_0(k,\,\varphi)$, and the differential cross section
in this limit takes form
\begin{eqnarray}
\frac{{\rm d}\sigma}{{\rm d}\varphi}=|f_0(k,\,\varphi)|^2=\frac{1}{4\pi k}\left(\frac{1}
{2\sin^2\left[\frac 12\left(\varphi+\frac{\eta\pi}{1-\eta}\right)\right]}+\frac{1}
{2\sin^2\left[\frac 12\left(\varphi-\frac{\eta\pi}{1-\eta}\right)\right]}\right.- \nonumber \\
\left.-\frac{\cos\left\{\left[2\frac{\Phi}{\Phi_0}-\left(2\left[\!\!\left[
\frac{\Phi}{\Phi_0}\right]\!\!\right]+1\right)\eta\right]\frac{\pi}{1-\eta}\right\}}
{\sin\left[\frac 12\left(\varphi+\frac{\eta\pi}{1-\eta}\right)\right]
\sin\left[\frac 12\left(\varphi-\frac{\eta\pi}{1-\eta}\right)\right]}\right).\label{eq35}
\end{eqnarray}
In particular, if the vortex flux is equal to an integer multiple of
a semifluxon, $\Phi=\frac n2\Phi_0$, then
\begin{equation}
\frac{{\rm d}\sigma}{{\rm d}\varphi}=\frac{1}{8\pi k}\frac{\sin^2\left(\frac{\eta \pi}{1-\eta}\right)}
{\left[\sin^2\left(\frac{1}{2}\frac{\eta\pi}{1-\eta}\right)-\sin^2\left(\frac \varphi 2\right)\right]^2}\,,\qquad
{\rm even} \,\,n,\label{eq36}
\end{equation}
\begin{equation}
\frac{{\rm d}\sigma}{{\rm d}\varphi}=\frac{1}{2\pi k}\frac{\sin^2\left(\frac{\varphi}{2}\right)\cos^2\left(
\frac 12\frac{\eta\pi}{1-\eta}\right)}
{\left[\sin^2\left(\frac{1}{2}\frac{\eta\pi}{1-\eta}\right)-\sin^2\left(\frac \varphi 2\right)\right]^2}\,,\qquad
{\rm odd}\,\, n. \label{eq37}
\end{equation}
Cross section (36) coincides with the cross section in the case of
zero flux, which was first obtained in \cite{Des}. Note that cross
section (37) vanishes in the forward direction, $\varphi=0$.
Due to the divergence of the differential cross section at
$\varphi=\pm \eta\pi(1-\eta)^{-1}$, the total cross section,
$$
\sigma_{\rm tot}=\int\limits_{0}^{2\pi}{\rm d}\varphi\frac{{\rm d}\sigma}{{\rm d}\varphi},
$$
is infinite in the long-wavelength limit.
The effects of the finite thickness of the vortex core become
important at shorter wavelengths of the scattered particle.
\section{Scattering of a short-wavelength particle}
The dependent on $r_{\rm c}$ part of the scattering amplitude,
$f_{\rm c}(k,\,\varphi)$ (30), is determined by the structure of the
vortex core, i.e. by the distribution of the gauge field strength
and the Higgs field inside the core; note that, unlike
$f_0(k,\,\varphi)$, $f_{\rm c}(k,\,\varphi)$ is smooth and
infinitely differentiable function of $\varphi$. Since our topic is
the Aharonov-Bohm effect, we would like to make the region of the
core to be inaccessible for the quantum-mechanical particle, and,
hence, we impose a boundary condition on the wave function at the
edge of the core. In the context of the Aharonov-Bohm effect the
Dirichlet boundary condition is mostly used,
\begin{equation}
\psi|_{r=r_{\rm c}}=0,\label{eq38}
\end{equation}
i.e. it is assumed that quantum-mechanical particles are perfectly
reffected from the vortex core.
If condition (38) is imposed, then one gets
\begin{equation}
\fl f_{\rm c}(k,\,\varphi)=-\exp[2{\rm i}k(r_{\rm c}-\xi_{\rm c})-{\rm i}\pi/4]
\sqrt{\frac{2}{\pi k}}\sum\limits_{n\in\mathbb{Z}}\exp[{\rm i}n(\varphi-\pi)-{\rm i}\alpha_n\pi]
\frac{J_{\alpha_n}(kr_{\rm c})}{H_{\alpha_n}^{(1)}(kr_{\rm c})}.\label{eq39}
\end{equation}
Unlike the case of $f_0(k,\,\varphi)$ (29), the infinite sum in (39)
cannot be calculated explicitly. However, the summation can be
performed in the case $kr_{\rm c}\gg 1$, yielding \cite{Si5}
\begin{eqnarray}
\fl f_{\rm c}(k,\,\varphi)=-\exp[2{\rm i}k(r_{\rm c}-\xi_{\rm c})](1-\eta)\sqrt{\frac{r_{\rm c}}{2}}
\sum\limits_{l}\sqrt{\cos[\frac 12(1-\eta)(\varphi-\pi+2l\pi)]}\times \nonumber \\
\times\exp\left\{{\rm i}\Phi\Phi_0^{-1}
(\varphi-\pi+2l\pi)-2{\rm i}kr_{\rm c}\cos[\frac 12(1-\eta)(\varphi-\pi+2l\pi)]\right\},\label{eq40}
\end{eqnarray}
where the finite sum is over integers $l$ that satisfy condition
(compare with (33))
\begin{equation}
-\frac{\varphi}{2\pi}-\frac 12\frac{\eta}{1-\eta}<l<-\frac{\varphi}{2\pi}+1+\frac 12 \frac{\eta}{1-\eta},\label{eq41}
\end{equation}
and terms of the order of $\sqrt{r_{\rm c}}O[(kr_{\rm c})^{-1/6}]$
and smaller are neglected.
Since $f_0(k,\,\varphi)$ is proportional to $k^{-1/2}$ (34), the
differential cross section in the short-wavelength limit, $kr_{\rm
c}\gg 1$, is given by
\begin{equation}
\frac{d\sigma}{d\varphi}=|f_{\rm c}(k,\,\varphi)|^2,\label{eq42}
\end{equation}
where $f_{\rm c}(k,\,\varphi)$ is given by (40).
In the case of a purely magnetic vortex ($\eta=0$) there is only one
term ($l=0$) in the sum in (40), and therefore the dependence on the
vortex flux disappears in the cross section
\begin{equation}
\frac{{\rm d}\sigma}{{\rm d}\varphi}=\frac 12r_{\rm c}\sin\frac \varphi 2 \qquad (0<\varphi<2\pi),\label{eq43}
\end{equation}
which is the cross section for scattering of a classical point
particle by an impenetrable cylindrical shell of radius $r_{\rm c}$,
see p.~1381 of \cite{Mor}. This result is easy to understand, since
the short-wavelength limit, $k\rightarrow \infty$, can be regarded
as the classical limit, $\hbar\rightarrow 0$, in view of relation
$k={\rm momentum}/\hbar$.
\begin{figure}
\includegraphics[width=260pt]{figure1}\\
\caption{Classical trajectories of scattered particles and scattering angle: \newline
$\omega_\eta=-\frac{\eta}{1-\eta}\pi$\, at $-\infty<\eta<0$,\,
$\omega_\eta=\left(2n-\frac{\eta}{1-\eta}\right)\pi$\, at
$\frac{2n-1}{2n}<\eta<\frac{2n}{2n+1}$\, ($n\geq 1$).}\label{1}
\end{figure}\normalsize
However, in the case of a vortex in conical space the dependence on
the vortex flux survives in the cross section in the
short-wavelength limit, if the number of terms in the sum in (40) is
more than 1. Before analyzing the situation with this number, let us
make a digression concerning scattering in classical mechanics.
\begin{figure}
\includegraphics[width=250pt]{figure2}\\
\caption{Classical trajectories of scattered particles and scattering angle: \newline
$\omega_\eta=\frac{\eta}{1-\eta}\pi$\, at $0<\eta<\frac 12$,\,
$\omega_\eta=\left(\frac{\eta}{1-\eta}-2n\right)\pi$\, at
$\frac{2n}{2n+1}<\eta<\frac{2n+1}{2n+2}$\, ($n\geq 1$).}\label{2}
\end{figure}\normalsize
If the vortex core is impenetrable for a classical point particle,
then its scattering does not depend on the vortex flux and is purely
kinematic, if the thickness of the vortex core is neglected. There
is no scattering in coordinates $\tilde{r}$, $\tilde{\varphi}$ (see
(17)), and, going over to the angular variable $\varphi$, one gets
classical trajectories depicted in figures 1 and 2, where the vortex
is directed perpendicular to the plane of the figure and its
position is indicated by the dot. The scattering angle is
independent of the impact parameter and is equal to $\omega_\eta$ or
$-\omega_\eta$ ($0\leq \omega_\eta\leq \pi$) depending on the side
from which the particle approaches the vortex. Depending on the
value of $\eta$, the trajectories either do not intersect (figure 1)
or do intersect (figure 2); the value of $\omega_\eta$ itself
depends on $\eta$. The region of angles
$-\omega_\eta<\varphi<\omega_\eta$ on figure 1 may be denoted as the
region of shadow (no objects from this region can be seen by an
observer to the left of the vortex). The region of angles
$-\omega_\eta<\varphi<\omega_\eta$ on figure 2 may be denoted as the
region of double image (every object from this region has double
image for an observer to the left of the vortex).
Returning to quantum-mechanical scattering, we note that the number
of terms in the distorted incident wave in (32) is equal to the
number of finite terms in the scattering amplitude in the
short-wavelength limit (40). This number denoted in the following by
$n_l$ is even in the region of classical shadow or classical double
image and is odd otherwise. Moreover, the value of $n_l$ outside the
shadow is larger by 1 than that in the shadow, whereas the value of
$n_l$ outside the double-image region is smaller by 1 than that in
the double-image region. To be more precise, in the case
$-\infty<\eta<0$ we have $n_l=1$ outside the shadow and $n_l=0$ in
the shadow (the main contribution is decreasing as $\sqrt{r_{\rm
c}}O[(kr_{\rm c})^{-1/6}]$); in the case $0<\eta<1/2$ we have
$n_l=1$ outside the double-image region and $n_l=2$ in the
double-image region; in the case $1/2<\eta<2/3$ we have $n_l=3$
outside the shadow and $n_l=2$ in the shadow; in the case
$2/3<\eta<3/4$ we have $n_l=3$ outside the double-image region and
$n_l=4$ in the double-image region; and so on with increasing values
of $n_l$. In particular, in the case $0<\eta<1/2$, which is most
relevant for the cosmic string phenomenology, we obtain the
differential cross section in the short-wavelength limit:
\begin{equation}
\frac{d\sigma}{d\varphi}=\frac 12r_{\rm c}(1-\eta)^2\cos[\frac 12(1-\eta)(\varphi-\pi)],\quad
\frac{\eta\pi}{1-\eta}<\varphi<2\pi-\frac{\eta\pi}{1-\eta},\label{eq44}
\end{equation}
and
\begin{eqnarray}
\fl \frac{d\sigma}{d\varphi}=r_{\rm c}(1-\eta)^2\Biggl\{\cos\left(\frac 12(1-\eta)\varphi\right)
\sin\left(\frac 12\eta\pi\right)+ \Biggr.\nonumber \\
\fl \Biggl.+\sqrt{\sin^2\!\left(\frac 12\eta\pi\right)\!-\!\sin^2\!\left(\frac 12(1\!-\!\eta)\varphi\right)}
\cos\!\left[2\pi\Phi\Phi_0^{-1}\!+\!4kr_{\rm c}\sin\!\left(\frac 12(1\!-\!\eta)\varphi\right)\cos\!\left(
\frac 12\eta\pi\right)\right]\Biggr\}, \nonumber \\
-\frac{\eta\pi}{1-\eta}<\varphi<\frac{\eta\pi}{1-\eta}.\label{eq45}
\end{eqnarray}
In the strictly forward direction, we get
\begin{equation}
\frac{d\sigma}{d\varphi}=2r_{\rm c}(1-\eta)^2\sin(\frac 12\eta\pi)\cos^2(\pi\Phi\Phi_0^{-1}),\quad \varphi=0.\label{eq46}
\end{equation}
In particular, if the vortex flux is equal to an integer multiple of
a semifluxon, $\Phi=\frac n2\Phi_0$, then (45) takes form
\begin{eqnarray}
\fl \frac{d\sigma}{d\varphi}=r_{\rm c}(1-\eta)^2\Biggl\{\cos\left(\frac 12(1-\eta)\varphi\right)
\sin\left(\frac 12\eta\pi\right)\pm \Biggr.\nonumber \\
\fl \pm\Biggl.\sqrt{\sin^2\left(\frac 12\eta\pi\right)-\sin^2\left(\frac 12(1-\eta)\varphi\right)}
\cos\left[4kr_{\rm c}\sin\left(\frac 12(1-\eta)\varphi\right)\cos\left(\frac 12\eta\pi\right)\right]\Biggr\}, \nonumber \\
-\frac{\eta\pi}{1-\eta}<\varphi<\frac{\eta\pi}{1-\eta},\label{eq47}
\end{eqnarray}
where the upper (lower) sign corresponds to even (odd) $n$.
It should be noted that the total cross section is independent of
the vortex flux and finite in the short-wavelength limit:
\begin{equation}
\sigma_{\rm tot}=2r_{\rm c}(1-\eta),\label{eq48}
\end{equation}
where the last relation is also valid in general case
$1>\eta>-\infty$; the decreasing terms of order $r_{\rm c}O[(kr_{\rm
c})^{-1/3}]$ and smaller are omitted in (48), as well as in
(44)-(47).
\section{Discussion}
Usually, the effects of non-Euclidean geometry are identified with
the effects which are due to the curvature of space. This is not the
case in general, and there are spaces which are flat almost
everywhere but non-Euclidean even in the vast region where they are
locally flat; this gives rise to non-Euclidean effects in such a
region. Conical space remains to be non-Euclidean in the whole even
if the transverse size of the region of nonzero curvature is shrunk
to zero. The non-Euclidean effect in this locally flat space is that
the space serves as a lens for propagating beams of light. Two
parallel beams after bypassing the axis of spatial symmetry from
different sides either converge (and intersect), or diverge,
depending on the value of deficit angle $2\pi\eta$. It is evident
that conical space serves as a concave lens in the case of negative
values of the deficit angle and as a convex lens in the case of some
bounded positive values of the deficit angle ($0<\eta<1/2$); it is
less evident, although is true, that, as the deficit angle grows
further ($1/2<\eta<1$), conical space serves in turn as a concave
and as a convex lenses, see figures 1 and 2.
Conical space per se is worth of interest, and, moreover, the
attention to this subject is augmented by the fact that conical
space emerges inevitably as an outer space of any topological defect
which is characterized by the nontrivial first homotopy group.
Although the ANO vortices yielding a noticeable amount of the
deficit angle have yet to be found, all hypothetical possibilities
in theory should be elaborated.
Bearing the above in mind, we consider the quantum-mechanical
scattering of a nonrelativistic particle by an ANO vortex. This is
the most general extension of the scattering Aharonov-Bohm effect to
the case when space outside a magnetic vortex is conical. From the
aspect of scattering theory, this corresponds to a situation when
the interaction with a scattering centre is not of the potential
type, see (25), and is even nondecreasing at large distances from
the centre, see (26). Such a fairly strong interaction violates the
conditions which are needed according to H\"{o}rmander \cite{Hor}
for constructing scattering theory in the case of the long-range
interaction. Despite this fact, a comprehensive scattering theory
has been constructed \cite{Si5}, and we rely mostly on the results
of this work.
The long-range nature of the interaction reveals itself in the
divergence of the scattering amplitude in the long-wavelength limit
in two directions which are symmetric with respect to the forward
direction, see (34); this should be compared with the case of
scattering by a magnetic vortex in Euclidean space, when the
amplitude diverges in one, forward, direction \cite{Aha}. The
long-range nature of the interaction is also revealed in the
distortion of the unity matrix in the relation between the
$S$-matrix and the scattering amplitude, see (27), and in the
distortion of the incident wave in the asymptotics of the scattering
wave function, see (32).
Both in the cases of Euclidean and conical spaces, the differential
cross section is a periodic function of the vortex flux with the
period equal to the London flux quantum. A peculiarity of conical
space is that the cross section vanishes in the forward direction,
if the vortex flux equals half of the London flux quantum, see (37)
at $\varphi=0$ for the long-wavelength limit and (46) at $\Phi=\frac
12\Phi_0$ for the short-wavelength limit.
In the present paper we have considered scattering of a
nonrelativistic spinless particle. For particles with spin, the
appropriate spin connections which are dependent on $\eta$ should be
introduced in hamiltonian (23). Thus, unlike scattering by an
impenetrable vortex in Eucledian space, such a scattering in conical
space depends on the spin of a scattered nonrelativistic particle.
In particular, for a spin -1/2 particle all results of the present
paper are modified in the following way: one should change
$\Phi\Phi_0^{-1}$ to $\Phi\Phi_0^{-1}\mp \frac 12 \eta$, where two
signs correspond to two spin states which are defined by projections
of spin on the vortex axis.
As the wavelength of a scattered particle increases, the effects of
the core structure of the vortex die out, and the differential cross
section becomes independent of the transverse size of the core, see
(35). Evidently, the long-wavelength limit corresponds to the
extremely quantum limit, when the wave aspects of matter are exposed
to the maximal extent.
As the wavelength of a scattered particle decreases, the effects of
the core structure of the vortex become prevailing. The
short-wavelength limit corresponds to the classical limit, when the
wave aspects of matter are suppressed in favour of the corpuscular
ones. Therefore, one would anticipate that, provided the vortex core
is made impenetrable for a scattered particle, the cross section
becomes independent of the vortex flux in this limit.
This anticipation is confirmed for the total cross section (48)
which corresponds to the classical expression that can be simply
interpreted as the quotient of the circumference of the core to the
half of the complete angle (note that only half of the core edge is
exposed to incident particles). However, this anticipation is
overturned for the differential cross section which remains to be
dependent on the vortex flux, see (45) for the case $0<\eta<1/2$.
For instance, if one considers a cosmic string corresponding to the
ANO vortex in the type-I superconductor, then the vortex radius is
equal to the correlation length, $r_{\rm c}=r_{\rm H}$, the vortex
flux is equal to an integer multiple of a semifluxon, $\Phi=\frac
n2\Phi_0$, and the differential cross section for the forward
scattering of a short-wavelength spinless particle by such a cosmic
string is (see (46) at small $\eta$)
\begin{equation}
\frac{d\sigma}{d\varphi}=4 \pi \, l_{\rm Pl}^2 \, r_{\rm H}^{-1}, \quad {\rm even} \, \, n , \label{eq49}
\end{equation}
\begin{equation}
\frac{d\sigma}{d\varphi}=0, \quad {\rm odd} \, \, n . \label{eq48}
\end{equation}
For a spin-1/2 particle the result is the same at even $n$ only, and
differs from zero otherwise:
\begin{equation}
\frac{d\sigma}{d\varphi}=16 \pi^3 \, l_{\rm Pl}^6 \, r_{\rm H}^{-5}, \quad {\rm odd} \, \, n . \label{eq51}
\end{equation}
This should be compared with the case $\eta=0$ when differential
cross section (43) vanishes in the forward direction, irrespective
of the value of the particle spin and the value of the vortex flux.
The reason of the discrepancy between the short-wavelength and
classical limits lies in expression (40) which yields nondecreasing
at $kr_{\rm c}\gg 1$ terms in the scattering amplitude; the vortex
flux is contained only in the phase of each term. The number of
these terms is determined by (41) and coincides with the number of
terms in the distorted incident wave in (32). This number is one at
$\eta=0$ and at $\eta<0$ (out of the region of classical shadow),
and, thence, the absolute value of the amplitude is independent of
the vortex flux in this case. The number is more than 1 at
$0<\eta<1$ (at $0<\eta<1/2$ in the region of classical double
image), and, thence, due to the interference of different terms, the
periodic dependence on the vortex flux survives in the
short-wavelength limit in the absolute value of the amplitude.
Although the given explanation is quite comprehensive, the simple
and transparent physical arguments, in our opinion, are lacking. The
Aharonov-Bohm effect, i.e. the periodic dependence on the flux of
the impenetrable vortex, is due to the nontrivial topology of space
with the excluded vortex region; this topology is the same both for
Euclidean space and for conical space with either positive or
negative deficit angle. Classical and quantum-mechanical motion
depends on the spatial geometry, i.e. on the value of the deficit
angle, and this is of no surprise. But a real puzzle is: Why conical
space with positive deficit angle distinguishes itself by the
discrepancy between the short-wavelength and classical limits?
\ack{}
Yu\,A\,S would like to thank the organizers of the AB-50 Workshop in
Tel Aviv University for kind hospitality during this extremely
interesting and inspiring meeting. The work was partially supported
by the Department of Physics and Astronomy of the National Academy
of Sciences of Ukraine under special program ``Fundamental
properties of physical systems in extremal conditions''.
\section*{References}
|
\section{Introduction} \label{sec:intro}
Many characteristics of flaring in active galactic nuclei (AGN) and
microquasars are best explained with relativistic motion of the
radiating regions in jets. Especially for microquasars the variability
is often best explained by variations in the structure of the
accretion disk and the jets; as discussed by many authors throughout
this volume, the connection between the disk and the large-scale jets
is well established even if the details are not yet fully known.
Our starting point for this report is the general model proposed for
both the microquasars and AGNs (see
refs.~\refcite{MirabelRodriguez1998,Marscher2002}, and references
therein). The connection between the accretion disk and the radio jet
in AGNs has been discussed from different angles, \emph{e.g.}, in
refs.~\refcite{Marscher2002}--\refcite{Chat08}, and in general the
scenario is believed to work along the following lines: part of the
accretion disk breaks off, and while the inner hot X-ray emitting
matter falls beyond the event horizon (causing a decrease in the X-ray
flux (see ref.~\refcite{Marscher2002} and references therein) the
outer parts get injected into the jet that is collimated and
accelerated before becoming visible at what is called the core of the
jet. There the jet plasma is believed to be compressed and heated by a
standing shock wave,\cite{Marscher2008,Chat09} causing a
multifrequency flare and a new knot becoming visible in the
jet.\cite{Savol02,Chat08} Apart from the jet core this is the same
process that was proposed for the microquasar GRS 1915+105 already
earlier.\cite{MirabelRodriguez1998} In this report we suggest an
additional component to the scenario.
Our study was motivated by the observations and idea of Miller-Jones
et al.,\cite{TwinPeaks} who found that certain double-peaked flares in
the Cygnus X-3 are best explained by treating the first flare as a
product of an outbreak of disk wind, later followed by a second flare
when a new jet element becomes visible further away in the jet. Here we study
the possibility that this model is at work also in AGNs and that -- in
addition to getting swallowed by the black hole or being injected into
the jet -- part of the collapsing accretion disk matter erupts from
the center as a wind-like outflow. We test if this outburst of matter --
originally having temperatures comparable to the inner parts of the
disk -- could produce observable effects or, in the most radical case,
cause a flare comparable to those happening in the jet, thus leading
to double-peaked flare. Modeling the causes of such
an event is beyond the scope of this early-phase report -- at this
point we focus on estimating what observational signatures a sudden
outflow of matter in the center of an AGN or microquasar could
produce.
In the following we describe our hypothesis for the first flare in the
center (Sec.~\ref{sec:thermal}) and discuss the expected signatures
and, in Sec.~\ref{sec:doubleflare}, discuss possible predictions of
the model concentrating on the application to AGNs. An example is
given in Sec.~\ref{sec:example} using BL Lacertae as a test case.
\section{Toy model for a thermal outburst in the center} \label{sec:thermal}
In this early phase of the study we use a very simplified model to see
if a thermal flare with reasonable parameters can, even in principle,
cause any of the observed features. For this test we take a
``spherical cow'' approach: we describe the radiating plasma as a
homogeneous sphere expanding at a constant speed and compute the
blackbody and the bremsstrahlung emission (taking the free-free
absorption and the light-travel time effects into account) at
different frequencies as a function of time.
For simplicity we take the matter to consist only of electrons and
positrons initially having a temperature comparable to that of the
accretion disk ($T_0\sim 10^{4-6}$ K for AGNs\cite{BonningEtAl}), and
calculate the matter density from the initial radius of the sphere
($r_S < R_0 \lesssim 3 r_{\rm S}$, where $r_{\rm S} = 2 GM/c^2$ is the
Schwartzschild radius of the black hole) and the mass of the
collapsing disk, estimated to be of the order of 1--10 $M_\odot$,
assuming the disk to contain mass accreted since the last major burst
--typically occurring once every couple of
years\cite{HovattaEtAl2007}-- at a rate of a few $M_\odot$/yr.
In the present model the sphere of plasma immediately starts to expand
with a constant speed --which we assume to be mildly relativistic, up
to 0.5 c (see, \textit{e.g.}, ref.~\refcite{Tombesi})-- and also cool
down due to adiabatic losses.
\section{Double flares -- observational features and predictions} \label{sec:doubleflare}
\paragraph{Different flare characteristics.}
As the mechanisms and the environments related to the first flare in
the center and the second flare at the core of the jet are different,
also the flares should differ from each other from spectral as well as
temporal points of view. Firstly, with typical AGN parameters the
sphere is optically thin for optical photons from the beginning,
making the flare bright in the optical waveband almost
immediately. Millimeter and radio flux, on the other hand, peak weeks
to months later and reach flux levels of only small fractions of those
of the optical peak. It is also important to notice that the
timescales or radiative signatures of first flare would not be
affected by similar Lorentz boosts or Doppler shifts than the flares
in the jet.
\paragraph{Single or double flares, depending on frequency.}
Whereas a thermal flare would most likely be observed and interpreted
as an optical flare without radio counterpart, a later flare (at the
radio core) is always expected to be bright throughout the
spectrum. This means that depending on the observing frequency, the
two flares would be observed either as a single- or double-peaked
flares (at the radio and optical wavelenghts, correspondingly).
\paragraph{Location of the core and delay between the flares.}
The delay between the two flares, $\Delta t$, would depend on the
distance of the jet core from the central black hole, on the speed
(and acceleration) of the jet, and on the light-travel-time
effects. Assuming, for simplicity, instant acceleration of the jet to
the speed corresponding to the apparent velocity $v_{\rm app}$ at
which the knots in the post-core jet (making an angle $\theta$ with
the line of sight) are seen to travel, the separation of the core is
simply $r_{\rm core} = \Delta t v_{\rm app} / \sin \theta$. Taking
into account the finite time required for the acceleration of the jet,
this may overestimate the actual distance.
\paragraph{Dip in the degree of polarization.}
Because the radiation from the first thermal flare is taken to be
unpolarized, and if the intensity of polarized emission is not
changed, the rise of the unpolarized radiation leads to a
corresponding decrease of the polarization. As a zeroth-order estimate
one can estimate the dip in the polarization simply from \( P = P_0 /
( 1 + I_{\rm flare}) \), where $P_0$ and $I_{\rm flare} = \Delta I /
I_0$ are the pre-flare polarization and the brightness increase
compared to the pre-flare level.
\paragraph{Flux levels start to rise before the second flare.}
If the first flare loads the jet with matter, this could lead to
the flux starting to rise already before the second flare. On one hand
this is due to the particles in the plasma being energized by the
strong magnetic turbulence near the center providing promising
conditions for efficient second-order Fermi acceleration,\cite{TD09}
leading to increased synchrotron emissivity. On the other hand, the
acceleration of the ''proto jet'' emitting region as a whole leads to
increased Lorentz boosting of the emitted radiation.
\paragraph{New ''knot'' in the jet after the second flare.}
Finally, when the jet-injected matter reaches the location of the VLBI
core of the radio jet, the standing shock (assumed to lie in the core)
compresses the plasma and accelerates particles on short timescales,
causing a multifrequency flare. After passing through the core, the
radiating plasma continues to travel along the jet and is now seen
as a bright knot.\cite{Savol02,Chat08} From this point on the ejected
plasma can be described using various shock-in-jet models.
\section{An example: BL Lacertae} \label{sec:example}
\begin{figure}[tH]
\centerline{\psfig{file=Tammi_2_f1.eps,width=0.8\linewidth}}
\vspace*{8pt}
\caption{Example lightcurves from a thermal flare in BL Lacertae for
optical (solid line) and three radio frequencies as given in the
plot. The dimmer and slower radio flares are only visible in the
small panel showing the micro-Jansky level peaks months after the
optical mJy-level flare. See text for details.\label{f1}}
\end{figure}
Let us illustrate the points raised in Sec.~\ref{sec:doubleflare} in
the case of the blazar BL Lac ($z=0.069$, $M \approx 10^8$ $M_\odot$).
By assuming the outbursting mass to be $M_{\rm in} = 10\, M_\odot$
(the major outbursts in the source happening on average every 2--3
years,\cite{NieppolaEtAl2009} leading to required mass accretion rate,
3--5 $M_\odot$/yr), the expansion speed $v_{\rm exp} = 0.4$ c, initial
radius $r_0 = 1.5\, r_{\rm S}$, initial temperature $T_0 = 10^5$ K and
adiabatic index $\gamma = 4/3$, we get an optical flare rising very
fast (within hours) to 8.5 mJy and declining in a few days with $S
\propto e^{-t/2}$, approximately, as shown in Fig.~\ref{f1}. At the
radio frequencies, however, the source becomes optically thin much
later and has a much lower peak flux density keeping the radio
signature of the flare invisible under the strong and variable
nonthermal flux from the jet. In essence this would be seen as an
optical flare without a radio counterpart.
To have even an order-of-magnitude reality check before detailed data
analysis, we compare this to observations of BL Lac made with with the
Kanata telescope in October
2008.\footnote{http://f.hatena.ne.jp/kanataobslog/20090106201125,
cited 1 Dec 2009.} Beginning on 11th of October 2008, we see an
optical flare rising from 8.4 mJy to 17.0 mJy within a few days, the
flux increasing by approximately 8.6 mJy. This is remarkably close to
the 8.5 mJy obtained from our zeroth-order model using first-guess
input parameters. For the duration of the flare the polarization
degree decreased from 18 \% to 11 \%. Using the same initial
polarization, we obtain the value of 9 \%, still interestingly near
the observed one. The comparison is, of course, very crude, but it is
sufficient to keep the proposed scenario plausible.
\section{Discussion and Conclusions} \label{sec:discussion}
In addition to the presented case for BL Lac, we have tested our toy
model for different AGNs and microquasars. The results (in
preparation) are promising: even with our very crude spherical-cow
model we have been able to reproduce lightcurves similar to those
observed both in AGN and microquasar objects. Some sources, however,
seem to be completely beyond the scope of the flux levels and
timescales obtainable with the present simple model. We are currently
collecting multifrequency data for detailed testing and further
development.
Further developments of the model will enable testing the X-ray
emission (enabling comparison with the observed anticorrelation
between the X-ray and radio fluxes \cite{Chat09}). Also the possible
role of nonthermal radiation, even in the case of a mostly thermal
flare, needs to be studied. Furthermore, the physicality of the model
will be improved by including different geometries and dynamics for
the outflow, as well as taking into account the presence of,
\emph{e.g.}, the dust torus for AGNs and companion star in
microquasar, as well as the differences in the density and composition
of the environment and the matter in the coronae of the black hole and
the accretion disk.
Finally we comment on two recent reports that are relevant to the
points presented in Sec.~\ref{sec:doubleflare}. Firstly, Marscher et
al.\cite{Marscher2008} reported an example of an optical double peak
with a single radio burst followed by a jet element in the jet of BL
Lacertae. The first flare was shown to happen in the acceleration and
collimation region before the radio core of the jet, following an
explosive event near the black hole; they explain the first optical
flare as a product of the increased radiation from the accelerating
plasma blob before the core. In our model the flare would take place
closer to the black hole -- otherwise our model assumes everything to
happen as described by Marscher et al.\cite{Marscher2008} The
published data are, however, still too sparse to either support or
disprove our model; further continuous and dense multifrequency
observations are needed to distinguish the possible non-jet flares
from those happening in the outflows or in the accretion
disk. Secondly, a very recent article by Villforth et
al.\cite{VillforthEtAl2009} published polarization data of the BL Lac
object OJ287. Their lightcurves included many peaks consisting mostly
of unpolarized flares. In this source the variability is considered to
be linked to the dynamics of a binary black hole with the secondary
black hole disrupting the accretion disk (see their paper for a review
of different models suggested for the source). Their results could
provide interesting and very useful tool in testing the proposed
model, as they emphasize the repeated double-peaked outbursts with the
first peak having no radio counterpart. Furthermore, they report dips
in the optical polarization during the burst. We acknowledge, however,
that the toy model presented here does not seem to be able to
reproduce the brightness of OJ287 flares. We dare not yet speculate
whether improved modeling of the black hole's environment could help,
or can this kind of an approach apply even in theory in an object
suspected to harbor a binary supermassive black hole
(ref.~\refcite{VillforthEtAl2009} and references therein).
To conclude, based on the preliminary results and within the limits of
currently available data we cannot rule out the possibility that in
some microquasars and AGNs certain flares can be due
to a thermal or partly thermal flare associated with the explosive
event that also dismisses parts of the accretion disk and injects
material into the jets. However, more data and improvements for the
model are needed to satisfyingly estimate the feasibility of
non-jet flares in these sources.
\section*{Acknowledgments}
We thank the Kanata telescope observation blog for their openness in
providing example data. Furthermore, Dr.~Tuomas Savolainen is
acknowledged for pointing out the possibility for using dip of the
polarization degree in testing the model.
|
\section{Introduction}
Some Quantum-Gravity models (see \cite{liv1,liv2,liv3,liv4,liv5} and references therein) postulate an inherent structure of spacetime (e.g. either a foamy, a discrete, or a lumpy spacetime) near the Planck scale $\lambda_{Pl}=\sqrt{G\hbar /c^3} \sim 10^{-35}m$. Since Lorentz symmetry is scale invariant (i.e. all scales are equivalent), the postulated existence of a special scale due to such QG effects can possibly lead to violations of Lorentz Invariance (LIV). One manifestation of such violations could be a dispersion in photon propagation, in which the speed of a photon in vacuo becomes dependent on its energy.
A detection and measurement of LIV effects would be indisputably invaluable to our understanding of the nature of spacetime at extremely small scales. Furthermore, even setting an upper limit on the magnitude of such effects can still prove very valuable, since measurements affected by physics at tiny Planck scales are rare and hard to perform, and can still be used to constrain and guide the relevant research. In this case, we have used high-quality measurements from GRB~090510 performed with both the GBM and LAT instruments on board the \textit{Fermi} observatory to set the most stringent limits to date on the magnitude of such energy-dispersion effects.
In the following, a brief introduction to the relevant formalism of LIV will be given, and some of the properties of GRB 090510 will be described. For more information on this study please refer to our Nature publication and its associated supplementary information \cite{nature}.
\subsubsection*{Lorentz-Invariance Violation}
Consider two photons of energies $E_h>E_l$ emitted simultaneously from a distant astrophysical source at redshift $z$. According to the postulated LIV effects these two photons will travel with different velocities and will arrive with a time delay $\Delta t$ equal to:
\begin{equation}
\label{eqDT}
$$ $\Delta t=s_{n}\frac{(1+n)}{2H_{0}}\frac{(E_{h}^{n}-E_{l}^{n})}{(M_{QG,n}c^{2})^{n}}\int_{0}^{z}\frac{(1+z')^{n}}{\sqrt{\Omega_{m}(1+z')^{3}+\Omega_{\Lambda}}}dz'$ $$
\end{equation}
Here $M_{QG}$ is the ``Quantum-Gravity mass'', a parameter that sets the energy scale at which such QG effects start to become important. Its value is assumed to be near the Planck Mass ($M_{Pl}=\hbar c/ \lambda_{Pl} \sim 10^{19}$GeV/$c^2$) and most likely smaller than it. The model-dependent parameter $n$ sets the order of the LIV and is assumed to be one or two, corresponding to linear ($\Delta t\propto \Delta E/M_{QG}$, with $\Delta E\equiv E_{h}-E_{l} \simeq E_{h}$) and quadratic ($\Delta t\propto \left(E_{h}/M_{QG}\right)^2$) LIV respectively. The model-dependent parameter $s_n$ is equal to plus or minus one, and it sets the type of LIV: a positive (negative) $s_n$ corresponds to a positive (negative) time delay or a speed retardation (acceleration) with an increasing photon energy.
In this analysis we used the times and the energies of the detected photons from GRB 090510 to set an upper limit on the strength of any LIV effects or equivalently a lower limit on $M_{QG}$. Since $M_{QG}$ is expected to be close and most likely smaller than the Planck Mass, a limit that is close to or, even better, over the Planck Mass is especially physically meaningful since it excludes most of the allowed parameter space, rendering any affected models highly implausible. It should be noted that not all QG models that predict a spacetime structure actually require such an energy dispersion. Instead, many of the models are consistent with LIV but they do not explicitly require it. A model that actually requires LIV, and can be directly constrained by such limits, is the stringy-foam model described in \cite{el08}.
\subsubsection*{GRB 090510}
Because of their short duration (typically with short substructure in the form of a series of narrow spikes) and cosmological distances, GRBs are well-suited for constraining LIV. In this study, we used measurements on the bright and short GRB 090510, which triggered both the LAT and GBM on May 10th 2009 at 00:22:59.97 UT (hereafter all times are measured relative to this trigger time). Ground-based optical follow-up spectroscopic data \cite{Rau2009}, taken 3.5 days later, exhibited prominent emission lines at a common redshift of $z=0.903\pm0.003$. The GBM light curve (figures \ref{fig:ET} b,c; 8 keV -- 40MeV) consisted of 7 main pulses. After the first dim short spike near trigger-time (hereafter called the ``precursor''), the flux went down to background levels. The main emission as detected by the GBM started at 0.53s and lasted for $\lesssim$0.5s. On the other hand, the main emission as detected by the LAT (figures \ref{fig:ET} a, d--f; E$\gtrsim$20MeV) started after the main GBM emission (0.65s vs 0.52s) and also extended to a significantly longer time scale (than the GBM emission) of about 200s. The emission detected by the LAT extended to an energy of about 31GeV, which is the highest energy ever detected from a short GRB. The fact that this 31GeV photon was detected shortly after the beginning of the burst ($\sim$0.8s), and the fact that the LAT-detected emission exhibited a series of very narrow spikes (up to few tens of ms width) that extended to high ($>$ tens of MeV) energies, allowed us to set stringent limits on the Quantum-Gravity mass. We used two independent methods to set these limits, described in the following section.
\section{Lower Limits on the Quantum-Gravity Mass}
\subsection{Using the arrival time of a single photon}
According to equation \ref{eqDT}, if we knew the LIV-induced time delay (${\Delta}t$) between two photons of energies $E_h>E_l$, then we should be able to make a measurement on $M_{QG}$ (for some model dependent parameters $s_n$ and $n$, and for a known measured $z$). However, this time delay cannot be measured because we do not know the exact emission times and because we cannot safely assume that the two photons were emitted from the same location (hence assume that there were no extra propagation delays due to non co-location). However, what we can do is to first assume that the higher-energy (HE) photon was emitted some time during a lower-energy (LE) emission episode (which starts from time $t_{start}$ and extends up to at least the HE-photon detection time $t_{HE}$), then calculate a \textit{maximum} time delay for the HE photon as ${\Delta}t_{max} = t_{HE}-t_{start}$, and finally calculate using equation \ref{eqDT} a \textit{lower limit} on the Quantum-Gravity mass.
Since the end time of the lower-energy emission episode cannot be safely assumed, the method mentioned above only works to constrain \textit{positive} time delays ($s_n=+1$). Also, this method is still valid even if the HE photon and the majority of the photons comprising the associated LE-emission episode are not emitted from the same exact location. Specifically, it only requires that there should be at least one LE photon that was emitted from the same location as the HE photon and was detected during the lower-energy emission episode, something that is generally safe to assume. Lastly, this method is only very weakly sensitive to any possible spectral lags occurring intrinsically at the GRB. Such lags have not been observed in short GRBs: they have been observed only in sub-MeV energies (and not in the higher-energy range of interest here), and even when they are observed (i.e. in long GRBs), they are of the order of a typical spike width (few tens of ms), which is considerably smaller than the time delays typically used in this analysis. As it will be shown, spectral lags, if they are actually present in GRB~090510, would have a negligible effect on our main results and would certainly not affect the conclusion of this study.
Setting stringent limits on the Quantum-Gravity mass requires an as-high-as-possible $E_h$ ($E_h$ usually ${\gg}E_l$ so $E_h-E_l{\simeq}E_h$) and an-as small-as-possible ${\Delta}t$. For GRB 090510, we used the highest-energy photon detected (31GeV) for settings such limits. Even if another lower-energy photon corresponded to a more stringent limit, to be conservative, we only used the 31GeV photon since its larger time delay makes it less sensitive to uncertainties in the choice of $t_{start}$ and to any intrinsic spectral lags. Furthermore, its higher energy renders it less probable of being a background event than an alternative lower-energy candidate. In the following calculations, again to be conservative, we used values for the GRB's redshift and for the 31GeV photon's energy reduced by one standard deviation.
The 31GeV photon had a reconstructed energy of $30.53^{36.32}_{27.97}$GeV ($1\sigma$ confidence intervals), it was detected 0.829s post-trigger, and it coincided with one of the GBM pulses. Very thorough analyses confirmed this event to be a real photon (instead of a background cosmic ray) associated with this GRB (e.g. from the diffuse galactic or extra-galactic emission). The properties of its associated event (absence of energy deposition in the anti-coincidence detector, the signature of an electromagnetic shower in the instrument, etc.) and the results from our event classification algorithms strongly supported its gamma-ray nature. Furthermore, its strong temporal (detected during the prompt emission) and directional (less than one PSF from the reconstructed by Swift direction) coincidence with the GRB supported its association with this source. And independently of these considerations, simple analysis regarding the rate of expected background events (both cosmic- and gamma-rays) resulted in a probability of less than $\sim10^{-7}$ ($>5\sigma$) of the background producing such an event.
The most conservative and of very high confidence assumption that can be made regarding the possible emission time of the 31GeV photon was that \textit{it was not emitted before the beginning of the precursor} (30ms before the trigger). For such an assumption, ${\Delta}t=0.829+0.03=0.859s$ and the associated lower limit on the QG mass for linear energy dispersion ($n=1$) is $M_{QG,1}\gtrsim1.19{\times}M_{Planck}$. An illustration that corresponds to this choice of $t_{start}$ is shown in sub-figure \ref{fig:ET}a with the black solid (n=1) and dashed (n=2) lines. It should be noted that the validity of this assumption depends on the absence of a second undetected precursor before the detected one. Even though the \textit{Fermi} LAT and GBM did not show evidence of any such second precursor, and even though no such two-precursor short GRBs have been detected, the existence of a second precursor starting at time $t_{pre,2}$ before the start of the first detected precursor $t_{pre,1}=-0.03s$ would decrease this most conservative limit by a factor of $\frac{t_{HE}-t_{pre,2}}{t_{HE}-t_{start}}$. The presence of any intrinsic spectral lags would correspond to a similar change in the limits. A negative spectral lag of $t_{lag} \sim 0.01s$ (in which the 31 GeV photon was emitted $t_{lag}$ time before the lower-energy associated events) would correspond to a decrease of our lower limit by a factor of $\frac{\Delta t + t_{lag}}{\Delta t} \simeq 1$ since here $t_{lag}<<\Delta t$.
Our most conservative limit, described above, can be improved by considering that it is actually very likely that the 31GeV photon is associated with only the main emission starting at $t_{start}=0.53s$ instead of with the whole burst (i.e. assumming that the 31GeV photon is far more likely to have been emitted during the main emission episode instead someting before it). This assumption is strongly supported by the fact that the 31GeV photon is expected to be emitted together with other lower-energy (sub MeV or MeV) photons that do not suffer from LIV delays and therefore mark its emission time. Using this assumption for the relevant emission time, the maximum time delay for the 31GeV photon becomes considerably smaller and our limits significantly stronger $M_{QG,1}{\gtrsim}3.42{\times}M_{Planck}$. Similarly, we can make somewhat less conservative assumptions regarding the lower-energy emission interval and associate the 31GeV photon with the start of the $>$100MeV emission or the start of the $>$1GeV emission, yielding stronger limits, yet with less confidence. Such limits correspond to smaller time delays, therefore they are relatively more sensitive to uncertainties in the choice of $t_{start}$ (because of now having to choose $t_{start}$ based on fewer photons) and to any possible intrinsic spectral lags. The results from all possible associations are shown in table I. As above, these limits are illustrated in sub-figure \ref{fig:ET}a. It should be noted that all the results of this method yield strong limits on $M_{QG}$ for linear LIV $n=1$, while they give considerably less constraining limits for the quadratic case. Also, as mentioned before, the above results are for the case of speed retardation $s_{n}=+1$ since we cannot safely assume the \textit{latest} emission time the 31GeV photon was emitted.
Finally, we note that the 31GeV photon arrives near the peak of a very bright and narrow spike in the soft gamma-ray lightcurve, which has a width of $\sim$10--20 ms (see figure \ref{fig:ET}). Such an association is not secure since the 31GeV photon could have landed over the spike just by chance.
However, just as an idea regarding the magnitude of the limits that could be attainable provided such an association was secure, we
calculate a significantly higher limit of $M_{QG,1}\gtrsim102{\times} M_{Planck}$ for both speed retardation and acceleration $s_{n}=\pm1$ (since now we can constrain both the earliest and latest possible emission time of the 31GeV photon). Similarly, a weaker
but independent and somewhat more robust limit on a possible negative time delay (which constrains the super-luminal case, $s_{n}=-1$) may be
obtained from the $\sim$0.75GeV photon that is observed during the precursor. This photon has a high probability
of being from GRB 090510 (a chance probability of the background producing such a photon of $\sim$1.2$\times 10^{-6}$ corresponding to $\sim 4.6 \sigma$), and the 1$\sigma$ confidence interval
for its energy is 693.6--854.4MeV. For the case of speed acceleration ($s_{n}=-1$), we can calculate an \textit{maximum} time delay of ${\Delta}t$<19ms and a limit of $M_{QG,1}\gtrsim1.33\times M_{Planck}$ (using again a reduced-by-one-standard-deviation photon energy in the calculation). These two associations with an individual spike are shown in figure \ref{fig:ET}a with the vertical dashed bands and in table I.
\subsection{Using the DisCan method}
An alternative method was also used for constraining any linear in energy LIV effects. This method, called DisCan \cite{DisCan} (Dispersion Cancellation), extracts dispersion information from \textit{all the LAT-detected photons} ($\sim$30 MeV -- $\sim$30 GeV) and \textit{does not involve binning in time or energy}. It is based on the fact that any QG-induced time delays would be expected to smear the spiky structure of the lightcurve. The DisCan method applies different trial spectral lags (time delays that are inversely proportional to the energy) to the lightcurve until it finds the one that maximizes a measure of its ``sharpness''. The trial spectral lag that accomplishes this is equal and opposite in sign to the sum of any QG-induced and intrinsic-to-the-GRB spectral lags.
The DisCan method was applied using the photons detected by the LAT during the interval 0.5-1.5s post-trigger, the burst interval with the most intense emission. The results were not significantly sensitive to any variations on the stop times for the interval ($\pm$0.25s) or the energy upper limits (1, 3 \& 100 GeV). The sharpness of the lightcurve was measured by using a cost function: the ``Shannon Information'' (eq. 11 of \cite{DisCan}). The value of this cost function is equal to the entropy (modulo a minus sign), and the difference of two values of the cost function are equal to the relative probability that one value is more likely for a given data set. Figure \ref{fig:DisCan1} shows the value of this cost function versus the trial spectral lag value. The minimum value of the cost function (most probable value) is $0^{+2}_{-18}$ ms/GeV. The errors here correspond to the trial spectral lag values that are 100 times less probable than the best value of 0ms/GeV, and are shown with the two vertical dashed lines in figure \ref{fig:DisCan1}.
\begin{figure}[t]
\centering
\includegraphics[width=80mm]{DisCan_1}
\caption{The value of the cost function (Shannon Information) versus the trial spectral lag value ($\theta$). The best
value of $\theta$, equal to zero, is annotated, and shown as a vertical solid line. The two dashed vertical lines left and right
of the best value represent the $\theta$ values which are 100 times less probable (-18 and +2 ms/GeV respectively) than the best $\theta$ value, \textit{for the given data set}. Thus the contained interval between the two dashed lines is an approximate error region,
but does not reflect statistical uncertainties.}
\label{fig:DisCan1}
\end{figure}
However, this result does not comprise a definite measurement of the spectral lag value; instead it is just one of the results that are compatible with the inherent uncertainties associated with our choice of time interval and energy range and with the limited statistics of the dataset. To estimate these uncertainties, a bootstrap analysis was performed, in which the DisCan method was applied to a randomized data set (a set produced by randomizing the association between the energies and times of the events). The reassignment destroyed any correlation with the energy, and therefore it removed any spectral lags. As a result, the application of this method to a randomized dataset should on average measure a spectral lag value that is equal to zero, while the width of the distribution of spectral lag values would be a measure of the statistical uncertainty of the measurement associated with our dataset. The results of the bootstrap analysis are shown in figure \ref{fig:DisCan2}, which shows the distribution of the measured spectral lags for each of the 100 randomized datasets. The measured spectral lags lie within a value of $<30ms/GeV$ for 99\% of the cases, and within a value of $<10ms/GeV$ for 90\% of the cases.
\begin{figure}[t]
\centering
\includegraphics[width=80mm]{DisCan_2}
\caption{For the interval analyzed in Figure \ref{fig:DisCan1}, to gauge uncertainty due to statistical variations
we generated 100 realizations with the photon times randomized. This figure shows the distribution of the minimum trial spectral-lag value $\theta_{min}$ for these 100 realizations. From this distribution, the 99\%CL for $\theta_{min}$ is 30ms/GeV.}
\label{fig:DisCan2}
\end{figure}
The combined result from minimizing the cost function for the actual GRB 090510 dataset and from the bootstrap analysis on the randomized datasets is an upper limit on the energy dispersion equal to $<$30 ms/GeV or $M_{QG}>1.22 \times M_{Planck}$ for linear energy dispersion of either sign $s_n=\pm 1$ at the 99\% confidence level. This result is shown in table I, along with the results from the previous method. For reasons similar to those advanced in the previous subsection (improbability of intrinsic lags or fortuitous cancellation of quantum gravity and intrinsic dispersion) we take this result as an upper limit on LIV-induced dispersion (i.e. we ignore any possible intrinsic spectral lags).
\begin{table*}[ht]
\label{table}
\begin{center}
\small{
\begin{tabular}{|c|c|c|c|c|c|c|c|c|c|}
\hline
& $t_{start}$ & Limit on $|\Delta t|$ & Reasoning for $t_{start}$ or method& $E_{l}$ & Valid for $s_{n}$ & Confidence& Limit on $M_{QG,1}$ & Limit on $M_{QG,2}$\tabularnewline
& (ms) & (ms) &used for setting the limits & (MeV) & & & ($M_{Planck}$) & ($10^{10}GeV/c^{2}$)\tabularnewline
\hline
\hline
(a) & -30 & < 859 & start of any $<$MeV emission & 0.1 & +1 & very high & > 1.19 & > 2.99\tabularnewline
\hline
(b) & 530 & < 299 & start of main$<$MeV emission & 0.1 & +1 & high & > 3.42 & >5.06\tabularnewline
\hline
(c) & 630 & < 199 & start of main$>$0.1 GeV emission & 100 & +1 & high & > 5.12 & > 6.20\tabularnewline
\hline
(d) & 730 & < 99 & start of main$>$1 GeV emission & 1000 & +1 & medium & > 10.0 & > 8.79\tabularnewline
\hline
(e) & -- & < 10 & association with$<$1 MeV spike & 0.1 & $\pm1$ & low & > 102 & > 27.7 \tabularnewline
\hline
(f) & -- & < 19 & if 0.75GeV $\gamma$-ray from 1st spike& & -1 & low & > 1.33 & > 0.54 \tabularnewline
\hline
(g) & \multicolumn{2}{c|}{$|\Delta t/\Delta E|<30ms/GeV$} & Lag analysis of all LAT photons& -- & $\pm1$ & very high & > 1.22 & --\tabularnewline
\hline
\end{tabular}
}
\caption{Lower limits on the Quantum-Gravity mass scale associated with possible Lorentz
Invariance Violation, that we can place from the lack of time delay in the arrival of high-energy
photons relative to low energy photons, from our observations of GRB 090510.}
\end{center}
\end{table*}
\section{Conclusion}
We have used high-quality measurements on GRB 090510 performed by the GBM and LAT instruments on board the \textit{Fermi} spacecraft to constrain tiny variations on the speed of light in vacuo that are linear or quadratic to its energy. We used two independent methods to obtain conservative and unprecedented upper limits on the magnitude of such speed variations. Our limits ($M_{QG,1}\gtrsim$few$\times M_{Planck}$) strongly disfavour any models predicting such linear-in-energy variations in the speed of light. Limits of such strength (over the Planck mass) are especially physically meaningful, since they exclude almost all of the allowed parameter space, rendering any affected models highly implausible. Any models predicting quadratic LIV are not significantly affected by our results. Our limits can be used to guide future research in Quantum Gravity and in general give insight on the nature of spacetime at minuscule Planck scales.
\begin{figure*}[t]
\label{fig:ET}
\centering
\includegraphics[width=135mm]{GRB090510_Energy-time}
\caption{Panel (a): energy vs. arrival time with respect to the GBM trigger time for the LAT photons that passed the transient event
selection (red) and the photons that passed the onboard $\gamma$-ray filter (cyan). The solid
and dashed curves are normalized to pass through the 31GeV photon and represent the relation between a photon's energy and arrival time
for linear (n=1) and quadratic (n=2) LIV, respectively, assuming it is emitted at $t_{start}=-30ms$ (black; first small GBM pulse onset), 530ms (red;
main $<MeV$ emission onset), 648ms (green; $>100MeV$ emission onset), 730ms (blue; >GeV emission onset). Photons emitted at $t_{start}$ would be located
along such a line due to LIV time delays. Panels (b)--(f): GBM and LAT lightcurves, from lowest to highest energies. Panel (f) also
overlays energy vs. arrival time for each photon, with the energy scale displayed on the right side. The dashed-dotted vertical lines show our 4
different possible choices for $t_{start}$. The gray shaded regions indicate the arrival time of the 31GeV photon (on the right) and of a
750MeV photon (during the first GBM pulse) (on the left), which can both constrain a negative time delay. Panels (b) and (c) show background
subtracted lightcurves for the GBM detectors. Panels (d)--(f) show, respectively, LAT events passing the onboard $\gamma$-ray filter, LAT transient class events with $E>100MeV$, and LAT transient class events with $E>1$ GeV. In all lightcurves, the time-bin width is 10ms.
}
\end{figure*}
\bigskip
\begin{acknowledgments}
The Fermi LAT Collaboration acknowledges support from a number of agencies and institutes
for both the development and the operation of the LAT as well as scientific data analysis.
These include NASA and DOE in the United States, CEA/Irfu and IN2P3/CNRS in France, ASI and
INFN in Italy, MEXT, KEK, and JAXA in Japan, and the K. A. Wallenberg Foundation, the Swedish
Research Council and the National Space Board in Sweden. Additional support from INAF in Italy
for science analysis during the operations phase is also gratefully acknowledged.
The Fermi GBM Collaboration acknowledges the support of NASA in the United States and DRL
in Germany.
\end{acknowledgments}
\bigskip
|
\section{Introduction}
\def{\mathcal B}_{\mathrm{cc}}{{\mathcal B}_{\mathrm{cc}}}
In this paper, we will apply methods of two-dimensional
sigma-models to two different problems in four-dimensional gauge
theory.
The first problem involves the relation of gauge theory to quantum
integrable systems. Vacua of massive two-dimensional gauge
theories with ${\mathcal N}=2$ supersymmetry correspond unexpectedly
\cite{MNS,GS,GS2}
to the quantum eigenstates of a
quantum integrable system. This correspondence has recently been extended
\cite{NS,NSmore} to a much
wider class of examples. In the present paper, we will approach
this rather surprising relation from a new angle. We focus on
what is perhaps the most challenging example of the correspondence
in question. This is the case \cite{NS2} that the two-dimensional
theory arises by reducing a four-dimensional ${\mathcal N}=2$ supersymmetric
gauge theory to two dimensions with the help of the $\Omega$
deformation \cite{Nek}. The resulting theories are associated to
quantum integrable systems that arise by quantizing a
finite-dimensional classical phase space. By contrast, many purely
two-dimensional examples
are more directly described in terms of quantum
spin systems.
Our basic idea is to map this problem to a brane construction in
two dimensions. Under certain conditions, a two-dimensional
$A$-model admits unusual branes \cite{KO} whose existence brings
noncommutativity into the $A$-model in several related ways
\cite{Kap,AZ,KW,GW}. Of most direct relevance to us is an
$A$-brane construction that leads to quantization of
finite-dimensional classical phase spaces \cite{GW}. In this
construction, integrability is natural. The construction involves
a pair of branes; one is an ordinary Lagrangian $A$-brane, and the
second, which has been called the canonical coisotropic $A$-brane
${\mathcal B}_{\mathrm{cc}}$, is the most simple example of the unusual $A$-branes
introduced in \cite{KO}.
In section 2, we describe some background that may be helpful. We
recall the basic reason that massive supersymmetric gauge theories
in two dimensions are related to integrability. We review a
variety of facts about two-dimensional sigma-models and their
relation to four-dimensional gauge theories, including the brane
construction \cite{GW} that will be our main tool.
In section 3, we describe how to study the
$\Omega$-deformation via this framework. Here, as in \cite{NS2},
we consider the $\Omega$-deformation of ${\mathcal N}=2$ super Yang-Mills
theory on ${\mathbb R}^4$, defined by a $U(1)$ action that leaves fixed a
two-plane ${\mathbb R}^2\subset{\mathbb R}^4$. There is an immediate problem, as it
appears that both the $\Omega$-deformed Lagrangian and the
supersymmetry preserved by the $\Omega$-deformation are not what
we need to make contact with the brane construction of \cite{GW}.
To overcome this difficulty, we give a new interpretation of the
$\Omega$-deformation. The $\Omega$-deformation is defined using a
vector field $V$ that generates a $U(1)$ symmetry of spacetime. As
originally defined in \cite{Nek}, building on \cite{MNS,LNS1,LNS2}, the $\Omega$ deformation involves a
deformation of the Lagrangian that preserves part of the
supersymmetry. We will give an alternative description of the
$\Omega$-deformation that is valid (for our purposes, which do not
depend on the precise choice of a $U(1)$-invariant metric) away
from the zeroes of $V$: by a change of field variables, one can
remove the deformation from the Lagrangian while rotating the
unbroken supersymmetry.
Taking this into account, we show that $\Omega$-deformed
supersymmetric gauge theory, in the situation considered in
\cite{NS2}, reduces naturally to an $A$-model in two dimensions,
with precisely the brane setup of \cite{GW}. The most unusual part
of this construction is the exotic $A$-brane ${\mathcal B}_{\mathrm{cc}}$. It arises in
giving a two-dimensional interpretation to what in four dimensions
are simply the $U(1)$ fixed points.
Section 4 is devoted to applying our framework to a very
different-sounding problem. Here our aim is to make contact with
remarkable results \cite{AGT} linking the $\Omega$-deformation in
four dimensions with Liouville theory (and its higher rank
analogs) in two dimensions. Our method can be applied to this
situation in an interesting way and answers some of the questions,
although many points are not yet clear. In this application, the
important branes are all rotated or dual versions of the
coisotropic brane ${\mathcal B}_{\mathrm{cc}}$.
Though our discussion is applicable to any ${\mathcal N}=2$ supersymmetric
gauge theory, we often specialize to a convenient and large class
of such theories which arise by compactification of the
six-dimensional $(0,2)$ model of type $G$ (here $G$ is a simple
and simply-laced Lie group) on a Riemann surface $C$. For every
choice of a system of $A$-cycles on $C$, one gets in four
dimensions a gauge theory realization \cite{DG} in which the gauge
group is a product of copies of $G$. We call these generalized
quiver theories. Compactification on a two-torus reduces a
generalized quiver theory to a sigma-model in two dimensions in
which the target space is ${\mathcal M}_H$, the moduli space of Higgs
bundles on $C$, endowed with a hyper-Kahler metric \cite{Hitchin}.
To keep things simple, a number of calculations in the body of
this paper are carried out only for compactification on a
rectangular two-torus and only for special values of the
parameters of the $\Omega$-deformation. A more complete treatment
is given in the appendix.
Our strategy throughout this paper is qualitatively similar to
many applications of toric geometry in string theory. For
example, see \cite{CN}. Also, the brane construction we use in
section 3 to obtain an eigenvalue problem for the commuting
Hamiltonians of a quantum integrable system is a cousin of a
construction that has been analyzed in the literature on geometric
Langlands \cite{F1,FFR,F2,F3}. Our (limited) understanding of the
relation is described in section \ref{eigand}. Finally, our
results in section 4 are qualitatively in agreement with previous
arguments suggesting a relation between duality of Liouville
theory and what is often called quantum geometric Langlands
\cite{FF,F,Teschner,Teschner2}. We hope it will prove possible to
make this connection more precise.
\def{\mathcal L}{\mathcal L}
\def\widetilde W{\widetilde W}
\section{Some Background}\label{aspects}
The present section is devoted to describing some background that
may be helpful. None of these results are new.
In section \ref{linko}, we will review the basic reason that there
is a link between two-dimensional massive supersymmetric gauge
theories and integrability. In section \ref{realsections}, we
discuss some generalities about what it might mean to quantize a
complex integrable system. In section \ref{coiso}, we review some
relevant facts about the two-dimensional topological $A$-model and
recall how $A$-branes can be used for quantization. In section
\ref{little}, we describe some pertinent differential geometry.
\subsection{The Basic Link Between Gauge Theory And
Integrability}\label{linko}
\def{\mathrm d}{{\mathrm d}}
\def\mathrm{\mathrm}
For illustrative purposes, we consider a two-dimensional ${\mathcal N}=2$
supersymmetric gauge theory with $(2,2)$ supersymmetry and a gauge
group $U(1)^r$. (The general case of a gauge group $G$ of rank
$r$ can be treated similarly; it leads to a nonabelian but equally
tractable version of the Lagrangian (\ref{kerf}).)
The $U(1)^r$ gauge theory has $r$ vector multiplets. Their
gauge-invariant content can be described by the twisted chiral
multiplets
\begin{equation}\label{twch}\Sigma_a=\sigma_a-i\sqrt 2\theta^+\overline\lambda_{a\,+}-
i\sqrt 2\overline\theta^-
\lambda_{a\,-}+\sqrt 2\theta^+\overline\theta^-(D_a-i\star F_a)+\dots,~a=1,\dots,r\end{equation}
where $\theta^\pm$ and $\overline\theta^\pm$ are superspace
coordinates, $\sigma_a$ and $\lambda_a$ are ordinary scalar and
fermi fields, $D_a$ are auxiliary fields,
$F_a$ is the field strength of the $a^{th}$ $U(1)$ gauge field, and $\star$ is the Hodge
star operator. We are interested in theories in which all vacua are massive, that is, admit only excitations
with positive mass. For this to be the case, any chiral multiplets that are present in the theory
must be massive; they can be integrated out, possibly making contributions to the effective twisted chiral superpotential
$\widetilde W(\Sigma_1,\dots,\Sigma_r)$ of the vector multiplets. The theory is then massive if $\widetilde W$
is sufficiently generic.
If so, the only important part of the effective action at low energies is the contribution of $\widetilde W$:
\begin{equation}\label{kerf}I_{\widetilde W}=\frac{1}{2\sqrt 2}\int d^2x\left({\mathrm d}\theta^+{\mathrm d}\overline\theta^-\widetilde W(\Sigma_1,\dots,
\Sigma_r)+\mathrm{c.c.}\right).\end{equation}
This is the action of a topological field theory; by dropping
higher derivative terms in the action, we have effectively taken
the masses of all excitations to infinity. The quantum states
obtained by quantizing (\ref{kerf}) are the same as the vacua in
the infinite volume limit of an underlying physical theory whose
action consists of $I_{\widetilde W}$ plus irrelevant terms of higher
dimension.
After performing the $\theta$ integrals, $I_{\widetilde W}$ gives for
the fermions $\lambda_a$ a mass matrix $m_{ab}=\partial^2\widetilde
W/\partial\sigma^a\partial\sigma^b$. If this is nondegenerate
(invertible) in every vacuum, as we will assume, then the fermions
are ``massive'' (but nonpropagating in the approximation of
(\ref{kerf}), as they have no kinetic energy). Let us look more
closely at the bosonic part of $I_{\widetilde W}$, which turns out to
be
\begin{equation}\label{zerf}I_{\widetilde W,B}=\int {\mathrm d}^2x \,\sum_a\left( D_a\mathrm{Re}\frac{\partial \widetilde W_a}
{\partial\sigma_a}+\star F_a \mathrm{Im}\frac{\partial\widetilde
W_a}{\partial\sigma_a}\right).\end{equation} Since the part of
$I_{\widetilde W,B}$ that involves the auxiliary fields $D_a$ has no
derivatives, the effect of the terms involving $D_a$ is simply to
impose constraints:
\begin{equation}\label{constraint}\mathrm{Re}\,\frac{\partial\widetilde W}{\partial\sigma_a}=0.\end{equation}
Let $\cmmib P$ be the locus (in a copy of ${\Bbb C}^r$ parametrized by
the $\sigma_a$) defined by these constraints. If $\widetilde W$ is
sufficiently generic, then $p_a=\mathrm{Im}\,\partial \widetilde
W/\partial\sigma_a$ is a good system of coordinates on $\cmmib P$.
We will analyze the dynamics of the fields $p_a$ and the $U(1)$
gauge fields $A_a$ on a two-manifold ${\mathbb R}\times S^1$, with metric
${\mathrm d} s^2={\mathrm d} t^2-{\mathrm d} x^2$, where $t$ is real-valued and $x\cong x+R$
(for some constant $R$) parametrizes $S^1$. We work in the gauge
in which the $t$ components of all gauge fields vanish, that is
$A_{t,a}=0$. The Gauss law constraint $\delta I_{\widetilde
W,B}/\delta A_{t,a}=0$ is $\partial p_a/\partial\sigma=0$, so in
this gauge, the $p_a$ are functions of $t$ only. We can
further fix the gauge so that the spatial parts of the gauge
fields are constants: $A_{x,a}=\phi_a/R$, where the $\phi_a$ are
angular variables that depend only on $t$. The action then reduces
to
\begin{equation}\label{polly}I'=\int {\mathrm d} t\sum_a p_a\frac{{\mathrm d}\phi_a}{{\mathrm d} t}.\end{equation}
This is an integrable system written in action-angle coordinates.
The $p_a$ are the action variables and the $\phi_a$ are the angle
variables. The nonzero Poisson brackets are
$\{p_a,\phi_b\}=\delta_{ab}$. In particular, the $p_a$ are
Poisson-commuting and quantum states can be labeled by the values
of the $p_a$.
However, not all values of the $p_a$ occur. The reason for this
is that the space $T$ obtained by specifying the values of the
$p_a$ is compact; it is a torus, parametrized by the $\phi_a$. In
the WKB approximation, the good values of the $p_a$ are those for
which the one-form $\omega=\sum_ap_a{\mathrm d}\phi_a$ has all its periods
integer multiples of $2\pi$. Each $\phi_a$ parametrizes a circle
$S_a$, and the basic periods are
$\int_{S_a}{\mathrm d}\phi_b=2\pi\delta_{ab}$. So the condition for
$\omega$ to have all integer periods is simply that the $p_a$
should be integers. Recalling the definition of the $p_a$, our
conclusion is that
\begin{equation}\label{polyp}\mathrm{Im}\,\frac{\partial\widetilde W}{\partial\sigma_a}=n_a,~n_a\in{\Bbb Z}.\end{equation}
Evidently, this result can be combined with eqn.
(\ref{constraint}) to give the holomorphic relation
\begin{equation}\label{zolyp}\frac{\partial\widetilde W}{\partial\sigma_a}=in_a, ~n_a\in{\Bbb Z}\end{equation}
or equivalently
\begin{equation}\label{olyp}\exp\left(2\pi \frac{\partial\widetilde W}{\partial\sigma_a}\right)=1\end{equation}
which characterizes the quantum states of the topological field
theory with action $I_{\widetilde W}$, or equivalently the vacua of an
underlying massive theory whose action differs from $I_{\widetilde W}$
by irrelevant operators.
Apart from possibly orienting the reader to the results of
\cite{MNS,GS,GS2,NS,NSmore,NS2}, the reason that we have explained
these matters here is to draw a lesson that will be important for
our derivation in the rest of this paper. The constraint
(\ref{olyp}) that determines the vacuum states is holomorphic in
$\widetilde W$, but in the derivation the real and imaginary parts of
$\partial\widetilde W/\partial\sigma_a$ have gone their separate ways.
Since this happens just in the two-dimensional derivation, it will
hopefully come as no surprise when something similar happens in
deriving this story from four dimensions.
\subsubsection{Validity Of The WKB Approximation}\label{wkb}
There remain two points to clarify about this derivation. First,
we have presented (\ref{zolyp}) or (\ref{olyp}) as the result of a
WKB approximation, but actually in the context of supersymmetric
gauge theory, these formulas are exact (modulo additive constants
that will be discussed). This is a standard result; for example,
see \cite{LNS} for a derivation via path integrals. The basic idea
of the derivation is to integrate over the gauge fields with the
action (\ref{zerf}). In the integral
\begin{equation}\label{kerfk}\int DA_a\,\exp\left(i\sum_a\int
F_a\,{\mathrm {Im}}\,\partial \widetilde W/\partial
\sigma_a\right)\end{equation} one would like to change variables
from $A_a$ to $F_a$, using the fact that $F_a$ contains almost the
full gauge-invariant content of $A_a$ and that (as the gauge group
is abelian) the volume of the space of gauge fields with given
curvature $F_a$ is independent of $F_a$. If the $F_a$ could be
treated as arbitrary two-forms, the resulting integral
\begin{equation}\label{omerf}\int DF_a\,\exp\left(i\sum_a\int F_a\,{\mathrm{Im}}\,\partial \widetilde W/\partial
\sigma_a\right)\end{equation} would simply give a delta function
setting ${\mathrm{Im}}\,\partial \widetilde W/\partial \sigma_a=0$.
Instead, $F_a$ is constrained by Dirac quantization, $\int
F_a/2\pi \in{\Bbb Z}$, which one can incorporate by including in the
path integral a factor $\sum_{n_a\in{\Bbb Z}}\exp(-i\sum_an_a\int F_a)$.
With this factor included, the integral over the $F_a$ now gives a
constraint ${\mathrm{Im}}\,\partial \widetilde W/\partial \sigma_a=
n_a$.
The result $\partial\widetilde W/\partial\sigma_a=i n_a$ has a more
elementary analog for chiral multiplets. If $\Phi_i=\phi_i+\theta
\psi_i+\theta^2F_i+\dots$ are chiral multiplets with
superpotential $W(\Phi_1,\dots,\Phi_s)$, then a vacuum is
characterized by $\partial
W(\phi_1,\dots,\phi_s)/\partial\phi_i=0$. No integer analogous to
$n_a$ enters because the auxiliary fields $F_i$ of the chiral
multiplet are independent complex fields. The result
(\ref{zolyp}) for twisted chiral multiplets of gauge theory is
slightly more complicated because the ``auxiliary field'' $\widetilde
F_a=D_a-i\star F_a$ in (\ref{twch}) is not an arbitrary complex
field; its imaginary part is subject to Dirac quantization.
In general, it would be unrealistic to expect the WKB
approximation to be always exact for an abstract integrable system
written in action-angle variables. Indeed, if $p_a,\phi_a$ are a
set of action-angle variables, then via a canonical
transformation, one can map them (in many different ways) to
another set of equally good action-angle variables $p'_a,\phi'_a$.
It is impossible for the WKB approximation to be valid in every
set of action-angle variables, so in general at best it will be
exact only for a system of action-angle variables that is in some
way distinguished. However, in the present context, the angle
variables $\phi_a$ are indeed distinguished, as they originate
from gauge fields; hence the conjugate variables $p_a$ are also
distinguished, modulo possible additive constants which will
indeed play a role. This makes it possible for the WKB
approximation to be exact.
The constraint (\ref{constraint}) is not holomorphic in $\widetilde W$,
but the condition defining the vacuum states of the underlying
gauge theory must be holomorphic. The WKB formula (\ref{polyp})
has indeed combined with the constraint (\ref{constraint}) to give
the desired holomorphy in (\ref{olyp}). A quantum correction to
(\ref{polyp}), apart from the possibility of adding a constant to
the right hand side, would spoil holomorphy. By adding a constant,
we mean replacing (\ref{polyp}) with
\begin{equation}\label{rufus}\mathrm{Im}\,\frac{\partial\widetilde W}{\partial\sigma_a}=n_a-\frac{\theta_a}{2\pi},
\end{equation}
for some angles $\theta_a$. The $\theta_a$ must be constants,
since if they were nontrivial functions of the $p_a$, this would
spoil holomorphy.
\def{\mathcal J}{{\mathcal J}}
For a perhaps fuller explanation, let us consider adding
$\theta$-angles to the underlying gauge theory action:
\begin{equation}\label{tangle}I_{\widetilde W}\to I_{\widetilde W}+\sum_a\theta_a\int\frac{F_a}{2\pi}.\end{equation}
In the above derivation, the one-form $\omega=\sum_ap_a\,{\mathrm d}\phi_a$
is then replaced by $\omega'=\sum_a(p_a+\theta_a/2\pi){\mathrm d}\phi_a$,
and this leads exactly to the generalized WKB condition
(\ref{rufus}). So this generalization is indeed something that we have to consider.
On the other hand, two-dimensional gauge theory is in general
invariant under $\theta_a\to \theta_a+2\pi m_a$. (This is usually
deduced from Dirac quantization of $\int F_a$, or alternatively
proved by observing that the unitary operator
$\exp(i\sum_am_a\oint_S A_a)$ brings about a shift
$\theta_a\to\theta_a+2\pi m_a$.) Allowing for arbitrary integers
$n_a$ in (\ref{rufus}) insures that the spectrum has this
invariance.
This discussion may raise the following question: is there a
renormalization of the angles $\theta_a$ in going from the
underlying Lagrangian to the exact condition (\ref{rufus}) that
characterizes the quantum states? In fact, this question is
equivalent to asking whether the effective superpotential $\widetilde
W$ has been identified correctly, since (as one can see by
comparing eqns. (\ref{zerf}) and (\ref{tangle})) introducing the
angles $\theta_a$ is equivalent to changing $\widetilde W$ to $\widetilde
W+i\sum_a \theta_a\sigma_a/2\pi$. So, given a microscopic theory, the problem
of finding the right constants in the WKB formula is part of the problem of correctly
computing the effective twisted chiral superpotential $\widetilde W$.
\subsubsection{Observables}\label{observables}
What are the natural observables of this integrable system? Going
back to the underlying gauge theory description, the natural
operators are those of the twisted chiral ring. The twisted
chiral ring of this theory is a polynomial ring generated by the
operators $\sigma_a$, $a=1,\dots,r$. They commute, and the
existence of these commuting operators can be regarded as an
explanation of why a detailed analysis of this system has led to
an integrable description in action-angle variables.
The derivation sketched above has led to a basis of quantum
states, characterized by (\ref{olyp}), in which the $\sigma_a$
have definite values. The chiral
ring generators -- and therefore all the operators of the chiral ring -- are diagonal in this
basis.
This was indeed the starting point of \cite{MNS,GS,GS2,NS,NSmore,
NS2}, where the conditions (\ref{olyp}) that determine the
eigenvalues of the chiral ring generators were interpreted in
terms of Bethe ansatz equations of an integrable system presented
not in action-angle variables but in some alternative and
physically interesting description. The relation between the two
pictures is not well understood, and we will not shed light on it
in the present paper. Rather, in this paper, our goal is to
understand the analog of the above two-dimensional derivation in
the context of $\Omega$-deformed theories in four dimensions.
\subsection{Complex Integrable Systems And Their Real
Sections}\label{realsections}
\def{\mathcal M}{{\mathcal M}}
We want to put this in the context of ${\mathcal N}=2$ supersymmetric gauge
theories in four dimensions. Let us denote as $\cmmib B$ the
Coulomb branch of the moduli space of vacua of such a theory.
(Depending on the spectrum of hypermultiplets, there may also be
Higgs branches of vacua or mixed branches, but they will not be
important in the present paper.) $\cmmib B$ parameterizes a family
of abelian varieties. We denote as $X$ the total space of this
fibration. $X$ is a complex symplectic manifold, with a
holomorphic symplectic form $\Omega$. The fibers of the fibration
$X\to\cmmib B$ are holomorphic submanifolds that are Lagrangian
with respect to $\Omega$. This means that $X$ is a completely
integrable Hamiltonian system in the complex sense \cite{DonW}.
The holomorphic functions on $\cmmib B$ are the action variables,
and the fiber coordinates are the angle variables.
An important class of examples \cite{DG} is derived by
compactification of the six-dimensional $(0,2)$ theory of type $G$
on a Riemann surface $C$. In this case, $X$ is a moduli space of
Higgs bundles on $C$, with structure group $G$, and the fibration
$X\to\cmmib B$ is the Hitchin fibration. The integrable system is
that of Hitchin \cite{Hitchin} and we denote $X$ as ${\mathcal M}_H$. It
may be that all ${\mathcal N}=2$ supersymmetric gauge theories in four
dimensions are related to Hitchin systems with suitable
singularities included. At any rate, we will continue the
discussion with this important class of examples in mind. When
formulated as four-dimensional gauge theories, these theories have
a generalized quiver structure.
In what sense might one quantize the completely integrable system
${\mathcal M}_H$? An important property of ${\mathcal M}_H$ is that it is birational
to a cotangent bundle $T^*{\mathcal M}$, where ${\mathcal M}$ is the moduli space of
stable holomorphic $G$-bundles on $C$. Such a ``birational''
equivalence holds after deleting complex submanifolds on both
sides. The holomorphic symplectic form of ${\mathcal M}_H$ maps to the
standard one on $T^*{\mathcal M}$.
This equivalence means that one can hope to map the commuting
Hamiltonians of the integrable system -- in other words, the
polynomial functions on $\cmmib B$ -- to holomorphic differential
operators on ${\mathcal M}$. This was first done for rank 2 Higgs bundles
(in other words, for the case that the gauge group of the
generalized quiver is a product of $SU(2)$'s) in \cite{HitchA} and
in generality in \cite{BD}; these constructions were based on
conformal field theory on $C$. An alternative argument starting
from four-dimensional gauge theory was explained at the end of
section 11.1 of \cite{KW}. An important fact is that the
differential operators in question act not on functions but on
sections of $K^{1/2}$, where $K$ is the canonical line bundle of
${\mathcal M}$. They commute with each other, just like the underlying
classical Hamiltonians.
\def{\mathcal D}{{\mathcal D}}
\def{\mathcal B}{{\mathcal B}}
For illustration, let us suppose that ${\mathcal M}$ is of complex dimension
1, which happens for rank 2 if $C$ is a Riemann surface of genus 1
with one marked point. (This corresponds to an ${\mathcal N}=2$ system that
is known as the ${\mathcal N}=2^*$ theory, whose Coulomb branch was related
to the corresponding Hitchin system in \cite{DonW}. It is also the
example considered in \cite{NS2}.) Consider a holomorphic function
$f$ on ${\mathcal M}_H$ that when restricted to the cotangent bundle $T^*{\mathcal M}$
is quadratic on each fiber of the cotangent bundle. (Such a
function can be derived from a Beltrami differential on $C$; in
gauge theory with a product of $SU(2)$ gauge groups, it
corresponds to a linear combination of the usual order parameters
of the Coulomb branch.) Letting $z$ be a local complex coordinate
on ${\mathcal M}$, and $p$ a fiber coordinate, $f$ is of the form
$f(z,p)=a(z)p^2$. In deformation quantization, $p$ maps to ${\mathrm d}/{\mathrm d}
z$ so $f$ will map to a second order holomorphic differential
operator ${\mathcal D}_f$ on $C$. Picking a local trivialization of
$K^{1/2}$, ${\mathcal D}_f$ is concretely given by a formula
\begin{equation}\label{givefor}{\mathcal D}_f=a(z)\frac{{\mathrm d}^2}{{\mathrm d}
z^2}+b(z)\frac{{\mathrm d}}{{\mathrm d} z}+c(z),\end{equation} with local
holomorphic functions $a,b,c$. Here, $a$ can be read off from the
classical function $f=a(z)p^2$, but the usual quantum mechanical
problem of operator ordering affects $b$ and $c$. In principle,
$b$ and $c$ can be computed in $\sigma$-model perturbation theory
for the brane system described in section \ref{coiso}. It is a
non-trivial fact that unique $b$ and $c$ functions do exist such
that the operator ${\mathcal D}_f$ is a globally-defined holomorphic
differential operator acting on sections of $K^{1/2}$. (This is
not true if $K^{1/2}$ is replaced by any other line bundle.) For
various explanations, see \cite{HitchA}, \cite{BD}, or section 11
of \cite{KW}. The details are anyway not really pertinent to the
qualitative remarks we will make here.
This construction has been formally called
quantization, but it is not what physicists usually mean by
quantization. A holomorphic differential operator ${\mathcal D}_f$ is
constructed, but there is no Hilbert space that it acts on.
Indeed, what could such a Hilbert space be? ${\mathcal D}_f$ could act on
global holomorphic sections of $K^{1/2}$, but there are none.
${\mathcal D}_f$ could also act on global $C^\infty$ sections of $K^{1/2}$,
or on meromorphic sections of $K^{1/2}$ with prescribed poles or
with arbitrary poles. But none of these spaces is a Hilbert space
in a natural way.
\def{\mathcal H}{{\mathcal H}}
To get a Hilbert space, we should pick\footnote{The actual construction made
in \cite{NS2} is more complicated than we are about to explain. The reason is
that in their example, ${\mathcal M}_H$ is only birational to $T^*{\mathcal M}$, for a Riemann surface ${\mathcal M}$.
If one aproximates ${\mathcal M}_H$ as $T^*{\mathcal M}$, one introduces a bad point in ${\mathcal M}$ where the
Hamiltonians have poles. The curves
$\gamma$ considered in \cite{NS2} pass through the bad point and this causes
the definition of the Hilbert space to be more involved.} a closed curve $\gamma\in
{\mathcal M}$. $K^{1/2}$ restricts on $\gamma$ to the bundle of
complex-valued half-densities on $\gamma$, and these form a
natural Hilbert space ${\mathcal H}_\gamma$. ${\mathcal D}_f$ acts naturally on
${\mathcal H}_\gamma$ (as an unbounded operator that is densely defined) for
the following reason. Real-analytic sections of $K^{1/2}$ along
$\gamma$ form a dense subspace $W_\gamma\subset {\mathcal H}_\gamma$. By
definition, a real-analytic section $s$ of $K^{1/2}$ along
$\gamma$ can be extended to a holomorphic section of $K^{1/2}$ on
a small neighborhood of $\gamma$ in ${\mathcal M}$. ${\mathcal D}_f$ then acts
naturally on $s$, and the restriction of ${\mathcal D}_f s$ to $\gamma$ is
again in $W_\gamma$. So ${\mathcal D}_f$ acts naturally on the dense
subspace $W_\gamma$ of the Hilbert space ${\mathcal H}_\gamma$. Under
suitable hypotheses on $\gamma$, ${\mathcal D}_f$ will be an elliptic
operator with a discrete spectrum.
Thus, the function $f$ has been mapped to an operator ${\mathcal D}_f$ that
acts on a Hilbert space ${\mathcal H}_\gamma$. This is what we usually mean
by quantization, and it is what is meant by quantization in
\cite{NS2}. However, quantization in this physical sense is not a
property of the complex integrable system ${\mathcal M}_H$ alone. It depends
on a choice of the real curve $\gamma$. In fact, in \cite{NS2},
two different choices of $\gamma$ are considered. (We will derive
these two choices from gauge theory in section \ref{farend}.)
In an $n$-dimensional example, the analog of $\gamma$ is an
$n$-dimensional real subspace $N\subset {\mathcal M}$. (A subspace $N$ is
called real if letting $TN$ be its tangent bundle and $I$ the
complex structure of ${\mathcal M}$, one has $TN\cap I(TN)=0$. This implies
that ${\mathcal M}$ is a sort of complexification of $N$.) Let
$L\subset{\mathcal M}_H$ be a completion of $T^*N\subset T^*{\mathcal M}$. Hitchin's
system restricted to $L$ is an ordinary real integrable system,
and the construction above can be regarded as quantization in the
ordinary sense of this integrable system.
What we have described is a two-step process. Hitchin's
integrable system can be quantized at a formal level by the
construction of certain holomorphic differential operators. To get
an actual Hilbert space requires a choice of a real cycle
$N\subset{\mathcal M}$. We have described this for two reasons: to orient
the reader to the sense in which we aim to quantize ${\mathcal M}_H$, and
also to help motivate the two-step nature of the brane
construction to which we turn next. In the brane construction,
the formal quantization is associated with a single brane, the
coisotropic $A$-brane that in the context of the
$\Omega$-deformation we will call ${\mathcal B}_\varepsilon$. Construction
of an actual Hilbert space ${\mathcal H}$ depends on the choice of a second
brane, a Lagrangian $A$-brane ${\mathcal B}_L$.
\subsection{Coisotropic $A$-Branes}\label{coiso}
\def{\mathcal I}{{\mathcal I}}
The brane construction that we will need in this paper relies on
aspects of the two-dimensional topological $A$-model that are not
novel \cite{KO,Kap,AZ,KW,GW} but are perhaps also not well known.
We will here summarize the facts that will be used later in the
paper, without attempting full explanations.
Consider the $A$-model of a symplectic manifold $X$ with symplectic structure $\omega$.
An $A$-brane of the familiar sort is supported on a Lagrangian submanifold $L\subset X$ that is endowed with
a flat vector bundle. Such an $L$ automatically has half the dimension of $X$.
However, in general the $A$-model admits
additional branes known as ``coisotropic'' branes. (One reason that such branes are not
well known is that they do not arise for Calabi-Yau threefolds.) The support of such an $A$-brane is
a submanifold $Y\subset X$ whose dimension exceeds half of the dimension of $X$.
The most basic new case is the case that $Y=X$. This case will
suffice in the present paper. (The general case is a sort of
hybrid of this with the more familiar Lagrangian $A$-branes.)
Unlike Lagrangian $A$-branes, whose Chan-Paton bundle is flat, the
curvature $F$ of a coisotropic $A$-brane is necessarily nonzero.
Only the rank 1 case is understood; in this case $F$ is an
ordinary two-form. The condition for a rank 1 brane whose support
is precisely $X$ and whose Chan-Paton bundle has curvature $F$ to
be an $A$-brane is \cite{KO} that the linear transformation of the
tangent bundle defined by ${\mathcal I}=\omega^{-1}F$ should obey ${\mathcal I}^2=-1$.
It is then automatically true that ${\mathcal I}$ is an integrable complex
structure on $X$. The condition ${\mathcal I}^2=-1$ implies that $F$ is not
only nonzero but is in fact non-degenerate. To generalize this
to include a $B$-field, we simply replace $F$ by $F+B$.
Reading this construction backwards, $X$ is a complex manifold and
$\Omega=F+i\omega$ is a holomorphic $(2,0)$-form that is closed
and non-degenerate. Thus, $X$ is a complex symplectic manifold.
However, neither the complex structure of $X$ nor
$F=\mathrm{Re}\,\Omega$ are part of the definition of the $A$-model of
$X$; rather they were used to define a brane. Only
$\omega=\mathrm{Im}\,\Omega$ is used in defining the $A$-model of $X$.
The same $X$ may have many different structures of complex
symplectic manifold (not related to each other by exact
symplectomorphisms with respect to $\omega$, which are trivial in
the $A$-model) each with a holomorphic two-form whose imaginary
part is $\omega$ and whose real part is the curvature of some line
bundle. Each of these will lead to a different brane in the same
$A$-model of $X$. Concrete examples can be constructed in the
hyper-Kahler case of section \ref{hyperkahler}.
\def{{\cmmib N}}{{{\cmmib N}}}
\def{\mathcal B}{{\mathcal B}}
\def{\mathcal R}{{\mathcal R}}
\def{\mathbb R}{{\mathbb R}}
\def{\mathcal L}{{\mathcal L}}
Though not yet part of the standard toolkit of physicists,
coisotropic $A$-branes have very interesting properties. If ${\mathcal B}$
is a Lagrangian $A$-brane, supported on a Lagrangian submanifold
$L$ and endowed with a flat bundle $E$, then the space of
$({\mathcal B},{\mathcal B})$ strings is ordinarily finite-dimensional and not really
quantum mechanical in nature. Additively, it is the cohomology of
$L$ with values in the bundle $E\otimes E^*$ ($E^*$ is the dual
bundle to $E$).
By contrast, if ${\mathcal B}$ is a rank 1 coisotropic $A$-brane whose
support is $X$, then the space of $({\mathcal B},{\mathcal B})$ strings of fermion
number zero is additively the space of complex-valued functions on
$X$ that are holomorphic in complex structure ${\mathcal I}$; allowing the
fermion number to vary, the space of $({\mathcal B},{\mathcal B})$ strings is the
$\overline\partial$ cohomology of $X$. This is quite an unusual answer;
we are accustomed to holomorphic functions and $\overline\partial$
cohomology in the $B$-model, but not in the $A$-model. Of course,
the size of the space of $({\mathcal B},{\mathcal B})$ strings depends very much on
$X$. At one extreme, if $X$ is compact, the only global
holomorphic functions are constants. We will be interested in
cases in which $X$ admits many holomorphic functions. The extreme
case is the case that $X$ is an affine variety, admitting in a
sense as many global holomorphic functions as there are on ${\Bbb C}^n$.
(In our applications, $X$ will or will not have this property
depending on the precise choice of coisotropic $A$-brane.)
Even more remarkable is the ring structure that arises from the
joining of $({\mathcal B},{\mathcal B})$ strings (or equivalently, from the
multiplication of boundary vertex operators that represent such
strings). This ring is a noncommutative deformation of the ring
of holomorphic functions on $X$ (or its extension to include the
higher $\overline\partial$ cohomology). In the case that $X$ has lots
of holomorphic functions, the noncommutative ring ${\mathcal R}$ of
$({\mathcal B},{\mathcal B})$ strings is the ring that can be obtained by deformation
quantization of the ring of holomorphic functions with respect to
the holomorphic symplectic form $\Omega$. In other words, the
first order departure from commutativity is given by the Poisson
bracket $\{f,g\}=(\Omega^{-1})^{i j}
\partial_if\,\partial_jg$, with higher order corrections largely determined by associativity of the
operator product expansion. See \cite{Kap} or in more detail
\cite{KW}, section 11; see also \cite{Pestun} for a related
analysis.
\def{\mathcal O}{{\mathcal O}}
To understand the noncommutative structure of the ring of
$({\mathcal B},{\mathcal B})$ strings, recall that, in general, interactions of open
strings involve a noncommutative structure that is much more
complicated than deformation quantization of a finite-dimensional
manifold. It is exceptional to find a situation in which
noncommutativity survives but reduces to something as simple as
ordinary deformation quantization. This occurs in somewhat
similar ways in the presence of a strong $B$-field \cite{CDS,SW}
or in the $A$-model with a coisotropic brane; either the strong
$B$-field or the topological symmetry of the $A$-model can
eliminate most of the open string modes, reducing to a
finite-dimensional but still noncommutative story. Let us briefly
describe how this comes about.
Suppose that on the target space $X$ of a sigma-model we have a
metric $g$ and a $B$-field $B$. We consider a brane endowed with
Chan-Paton curvature $F$. We make no assumption in general that
$X$ is a complex manifold, and write $I,J=1,\dots,n$ for tangent
space indices to $X$. In sigma-model, one encounters \cite{SW} an
effective inverse metric
\begin{equation}\label{overdu}G^{IJ}=\left(\frac{1}{g+2\pi\alpha' (F+B)}g\frac{1}{g-2\pi\alpha'(F+ B)}\right)^{IJ}\end{equation}
and a noncommutativity parameter
\begin{equation}\label{nerdu}\theta^{IJ}=-(2\pi\alpha')^2\left(\frac{1}{g+2\pi\alpha'(F+
B)}(F+B) \frac{1}{g-2\pi\alpha' (F+B)B}\right)^{IJ}.\end{equation}
In sigma-model perturbation theory, in computing the operator
product expansion of boundary operators ${\mathcal O}_u$ and ${\mathcal O}_v$
associated to functions $u$ and $v$ on $X$, one meets symmetric
contractions proportional to $G^{IJ}\partial_Iv
\partial_Jv$, and antisymmetric contractions proportional to
$\theta^{IJ}\partial_Iu\partial_Jv$. For the sigma-model to
reduce to something as simple as deformation quantization, the
symmetric contraction must be negligible compared to the
antisymmetric one. The most familiar way to achieve this result
is to take $F+B\to\infty$, noting that $G^{IJ}\sim 1/(F+B)^2$
while $\theta^{IJ}\sim 1/(F+B)$. However, in the $A$-model, there
is another way to suppress the symmetric contractions. In the
context of the coisotropic $A$-brane ${\mathcal B}$ just described, $u$ and
$v$ are required to be holomorphic functions in complex structure
${\mathcal I}$. Moreover, $g$ is of type $(1,1)$ while $F+B$ is of type
$(2,0)\oplus (0,2)$. In this case, the symmetric contraction
vanishes, and the antisymmetric contraction is governed by
$\theta^{-1}=-(F+B)^{-1}_{2,0}$, that is, the $(2,0)$ part of
$-(F+B)^{-1}$. In our above presentation, we took $B=0$,
$F=\mathrm{Re}\,\Omega$, so the anticommutativity parameter is
$-\Omega^{-1}$.
\def{\mathcal S}{{\mathcal S}}
Now the question arises of whether we can use this framework to
see quantum mechanics, and not simply deformation quantization. To
do so, we need a Hilbert space on which the noncommutative ring
${\mathcal R}$ acts. For this, we consider a pair of $A$-branes -- a
coisotropic $A$-brane ${\mathcal B}$ of support $X$ and a rank 1 Lagrangian
$A$-brane ${\mathcal B}_L$ of support $L$. We write ${\mathcal L}$ and ${\mathcal S}$ for the
Chan-Paton bundles of ${\mathcal B}$ and ${\mathcal B}_L$, respectively. ${\mathcal L}$ and ${\mathcal S}$
are both endowed with unitary connections.
The curvature of ${\mathcal L}$ is a non-degenerate two-form $F$; the curvature of ${\mathcal S}$ vanishes.
Whatever $L$ is, the ring ${\mathcal R}$ of $({\mathcal B},{\mathcal B})$ strings will act on the space $\mathcal H$ of $({\mathcal B},{\mathcal B}_L)$
strings, by the usual operation of joining strings. The fact
that we must introduce a second brane to define $\mathcal H$ has
an obvious parallel with what we stated more naively in
section\footnote{In our presentation, we are eliding a few key
details that are described in \cite{GW}. Though $\mathcal H$ can
always be defined in the $A$-model, and has a Hilbert space structure
because it is the space of ground states of the sigma-model, this Hilbert space structure is
natural in the $A$-model only if $L$ is the fixed point set of an
antiholomorphic involution of $X$. And when this is the case,
the Hilbert space structure on $\mathcal H$ that is natural in the
$A$-model coincides with the naive one introduced in section
\ref{realsections} only in the semiclassical limit, that is, to
lowest order in sigma-model perturbation theory.}
\ref{realsections}.
Under certain conditions which we will now state, $\mathcal H$ can be interpreted in terms of
quantization of $L$.
The fact that $L$ is Lagrangian means by definition that $\omega$
vanishes when restricted to $L$. What about $F$? Let us consider
two contrasting cases. (There are also various intermediate cases,
but they will not be important for us.) If $F$ vanishes when
restricted to $L$, then $L$ is actually a complex
submanifold\footnote{For a proof of this statement, see section
\ref{little} -- though a different notation is used there with
$I$, $J$, and $K$ cyclically permuted. As will become clear in
this paper, it is difficult to find a single and uniformly
convenient notation.} in complex structure $I$. This case does not
lead to what physicists usually understand as quantization, but
can lead to interesting and purely holomorphic constructions of
spaces on which ${\mathcal R}$ acts, and it is important for geometric
Langlands \cite{KW}.
\def{\mathcal H}{{\mathcal H}}
The opposite case is the case that $F$ remains nondegenerate when
restricted to $L$. Thus, though Lagrangian with respect to
$\omega$, $L$ is symplectic with respect to $F$. If unitarity is
desired, one also requires that $X$ should have an antiholomorphic
involution (a symmetry of order 2) with $L$ as a component of its
fixed point set. Under these conditions, as found in \cite{AZ} in
examples and discussed more systematically in \cite{GW}, the space
${\mathcal H}$ of $({\mathcal B},{\mathcal B}_L)$ strings can be understood as a quantization of
$L$, with symplectic structure $F$ (and prequantum line bundle
${\mathcal L}\otimes {\mathcal S}^{-1}$, whose curvature is $F$). The basic reason
for this is easily explained. As usual, the physical states of
the $A$-model are the string ground states, which can be found by
quantizing the zero-modes of the string. In the case of the
$({\mathcal B},{\mathcal B}_L)$ strings, one finds that there are no fermion
zero-modes. The bosonic zero-modes describe the motion of the
string along $L$, and the relevant part of the action for these
modes is the Chan-Paton contribution. Writing $\cmmib A=\sum_i
p_i\, {\mathrm d} q^i$ for the connection on the prequantum line bundle
${\mathcal L}\otimes {\mathcal S}^{-1}$ ($p_i$ and $q^ j$ are a local system of
canonically conjugate coordinates on $L$ viewed as a symplectic
manifold with symplectic form $\omega_J$), the relevant action is
$\int{\mathrm d} t \,p_i \,{\mathrm d} q^ i/{\mathrm d} t$; quantization of the zero-modes
with this action is usually called quantization of $L$.
The functions on $L$ that can be most
naturally quantized as operators on ${\mathcal H}$ are
the functions that are restrictions to $L$ of holomorphic functions on $X$.
Such functions
are quantized by identifying them with $({\mathcal B},{\mathcal B})$ strings which then act naturally on ${\mathcal H}$.
If $X$ is an affine variety in complex structure ${\mathcal I}$, functions
on $L$ that are restrictions of holomorphic functions on $X$ are
dense in the space of functions on $L$. We consider two primary
applications in this paper. In section \ref{confblocks}, where we
discuss the relation \cite{AGT} between four-dimensional gauge
theory and two-dimensional Liouville theory, $X$ is indeed an
affine variety in complex structure ${\mathcal I}$. The holomorphic
functions on $X$ are numerous and act irreducibly in the
quantization. In section \ref{compom}, where we discuss the
relation \cite{NS2} of the $\Omega$-deformation to quantization of
Hitchin's integrable system, $X$ is not affine in complex
structure ${\mathcal I}$ (but instead is a fibration by abelian varieties
over an affine base $\cmmib B$). The holomorphic functions are the
commuting Hamiltonians of Hitchin's integrable system. Their
interpretation via $({\mathcal B},{\mathcal B})$ strings amounts to their
interpretation as commuting holomorphic differential operators on
${\mathcal M}$, as explained in detail in section 11 of \cite{KW}. To get a
Hilbert space ${\mathcal H}$ on which these operators can act, we need to
pick a real section of ${\mathcal M}$, as explained heuristically in section
\ref{realsections}. More fundamentally, we need to pick a
Lagrangian brane ${\mathcal B}_L$ in the $A$-model of ${\mathcal M}_H$ in symplectic
structure $\omega_K$.
Our two applications will involve different sides of the same
coin. But to explain this, we must now specialize to the
hyper-Kahler situation.
\subsubsection{The Hyper-Kahler Case}\label{hyperkahler}
In our examples, $X$ will be the Coulomb branch of the moduli
space of vacua of a four-dimensional theory with ${\mathcal N}=2$
supersymmetry, after compactification on a circle (or a two-torus)
to three or two dimensions. Thus $X$ will be hyper-Kahler. An
important class of examples \cite{DG}, already considered for
illustration above, are the generalized quiver theories in which
$X$ is a moduli space ${\mathcal M}_H$ of Higgs bundles on a Riemann surface
$C$.
Such an $X$ has a distinguished complex structure $I$ (in which
the Hitchin fibration is holomorphic) and another distinguished
complex structure $J$ (in which it parametrizes complex-valued
flat connections on $C$). We set $K=IJ$; $I$, $J$, and $K$ obey
the quaternion algebra. Together with a Riemannian metric $g$,
they define the hyper-Kahler structure of $X$. Any linear
combination ${\mathcal I}=aI+bJ+cK$, where $a,b,c$ are real and
$a^2+b^2+c^2=1$, is an integrable complex structure. This family
of complex structures is parametrized by $\mathbb{CP}^1\cong S^2$.
The three real symplectic forms, which are Kahler in complex
structures $I,J$, or $K$ respectively, are $\omega_I=gI$,
$\omega_J=gJ$, $\omega_K=gK$. Similarly, the three holomorphic
symplectic forms, which are of type $(2,0)$ with respect to $I,J$,
or $K$, are $\Omega_I=\omega_J+i\omega_K$, $\Omega_J
=\omega_K+i\omega_I$, and $\Omega_K=\omega_I+i\omega_J$. For
simplicity, we will assume the $B$-field to vanish.
We will be studying the $A$-model of $X$ in the symplectic
structure $\omega=\omega_K$. To define a coisotropic brane in this
situation, we can take $F=\omega_I\sin p + \omega_J\cos p$ for
some angle $p$. We must constrain $p$ and the hyper-Kahler metric
of $X$ so that $F/2\pi$ has integer periods and hence $F$ is the
curvature of some line bundle ${\mathcal L}\to X$. Given this, we get a
coisotropic $A$-brane with ${\mathcal I}=\omega^{-1}F=I\cos p -J\sin p $.
\def{\mathcal B}_{\mathrm{cc}}{{\mathcal B}_{\mathrm{cc}}}
For our eventual application, the most important case will be that
$p=0$, so $F=\omega_J$ and ${\mathcal I}=I$. (In examples arising by
compactification from ${\mathcal N}=2$ theories in four dimensions,
$\omega_J$ is cohomologically trivial, so there is no problem in
constructing a line bundle with curvature $F$.) The brane ${\mathcal B}$ so
obtained is a sufficiently basic example of a coisotropic
$A$-brane that it has been called the canonical coisotropic
$A$-brane ${\mathcal B}_{\mathrm{cc}}$. We have constructed it to be an $A$-brane in
the $A$-model of symplectic structure $\omega_K$. However, it has
additional supersymmetric properties and these will be important.
Suppose we take $\omega=\omega_I$; then $\omega^{-1}F=-K$, which
is again an integrable complex structure. So the same brane ${\mathcal B}_{\mathrm{cc}}$
is also an $A$-brane for another $A$-model, the one with
$\omega=\omega_I$. Finally, as $\omega_J$ is of type $(1,1)$ in
complex structure $J$, we see that ${\mathcal B}_{\mathrm{cc}}$ is a $B$-brane for the
$B$-model of complex structure $J$.
These three facts are related, in the following sense. The
topological supercharges of the $A$-model of symplectic structure
$\omega_I$, the $B$-model of complex structure $J$, and the
$A$-model of symplectic structure $\omega_K$ obey one linear
relation. They are linear combinations of two supercharges $Q$
and $Q'$ (this will be explained in detail in section
\ref{compom}). Any linear combination $uQ+vQ'$, with complex
coefficients $u,v$ that are not both zero, squares to zero and is
the topological supercharge of some topological field theory.
Varying the ratio $u/v$, this gives a family of topological field
theories, parametrized by $\mathbb{CP}^1$, which admit the same brane
${\mathcal B}_{\mathrm{cc}}$. For brevity, we will describe this by saying that ${\mathcal B}_{\mathrm{cc}}$
is a brane of type $(A,B,A)$.
As branes with multiple supersymmetric properties may be
unfamiliar, we will mention a much more obvious example that will
also be important in this paper. This is the brane ${\mathcal B}_*$ whose
support is all of $X$ and whose Chan-Paton line bundle ${\mathcal L}_*$ is
trivial. The support of ${\mathcal B}_*$ (being all of $X$) and the line
bundle ${\mathcal L}_*$ (being flat) are both holomorphic in every complex
structure ${\mathcal I}=aI+bJ+cK$. So the brane ${\mathcal B}_*$ is a $B$-brane in a
family of $B$-models parametrized by $\mathbb{CP}^1$. We summarize
this by saying that
${\mathcal B}_*$ is a brane of type $(B,B,B)$. Again, the various supercharges
are linear combinations of any two of them.
The differential geometry of branes with multiple supersymmetric
properties is further described in section \ref{little}.
\subsubsection{First Among Equals}\label{distinguished}
In this hyper-Kahler situation, we can define a plethora of
topological field theory structures -- the $B$-model in any
complex structure $aI+bJ+cK$, or the $A$-model in any symplectic
structure $a\omega_I+b\omega_J+c\omega_K$. However, in the
important case
\cite{DG} that $X$ is
actually a moduli space ${\mathcal M}_H$ of Higgs bundles on a Riemann
surface $C$, with some gauge group $G$, some of these structures
are more special than others.
\def{\mathcal A}{{\mathcal A}}
Almost all of these two-dimensional topological field theories
depend on the complex structure of $C$. But some do not. In one
of its complex structures, customarily called $J$, ${\mathcal M}_H$ is the
moduli space of flat connections on $C$ with values in the
complexification of $G$. This is a complex symplectic manifold in
a completely natural way, independent of any choice of metric or
even complex structure on $C$. Thus, both the complex structure
$J$ and the holomorphic two-form $\Omega_J=\omega_K+i\omega_I$ do
not depend on any property of $C$ beyond its orientation. In fact,
$\Omega_J$ can be defined by the formula
\begin{equation}\label{zork}\Omega_J=\frac{i}{4\pi}\int_C{\mathrm {Tr}}\,\,\delta {\mathcal A}\wedge\delta{\mathcal A},\end{equation}
where ${\mathcal A}=A+i\phi$ is a complex-valued flat connection; this
formula does not use a metric or complex structure on $C$, so it
makes clear the topological nature of $\Omega_J$.
Accordingly, the $B$-model of type $J$ and the $A$-models of types
$\omega_I$ and $\omega_K$ are special -- they do not depend on the
choice of a complex structure on $C$. In that sense, branes of
type $(A,B,A)$ are special, compared to say branes of type
$(B,B,B)$, $(B,A,A)$, or $(A,A,B)$, all of which have a mixture of
properties that do or do not depend on a complex structure on $C$.
As explained earlier, a brane on ${\mathcal M}_H$ with multiple
supersymmetric properties is a brane in a whole family of
topological field theories, parametrized by $\mathbb{CP}^1$. In
general, this family depends on the complex structure of the
underlying Riemann surface $C$. Precisely in the case of a brane
of type $(A,B,A)$, the relevant family of topological field
theories is independent of the complex structure of $C$. Our
applications are based on this family, as is also the gauge theory
approach to geometric Langlands \cite{KW}.
In a four-dimensional ${\mathcal N}=2$ theory with $U(1)_R$ symmetry, the
forms $\omega_J$ and $\omega_K$ are rotated by the $U(1)_R$
symmetry, which ensures that their cohomology classes (which would
have to be rotation-invariant) must vanish. This is actually true
even without $U(1)_R$ symmetry, as long as hypermultiplet bare
masses vanish. For generalized quiver theories associated to a
Riemann surface $C$ without marked points, this can be shown in
terms of differential geometry as follows.
Describing a Higgs bundle by a gauge
field $A$ with Higgs field $\phi$, the exactness of $\omega_J$ and
$\omega_K$ follows from an explicit formula
\begin{equation}\label{zobox}\omega_J+i\omega_K=\delta\left(\frac{1}{\pi}\int_C{\mathrm {Tr}}\,\phi_z\delta
A_{\overline z}\right).\end{equation} (In the presence of a
hypermultiplet bare mass, $\phi_z$ has a pole whose residue has an
eigenvalue proportional to the mass, and the above argument fails
because the one-form in parentheses is not gauge-invariant; that
is, it is not the pullback of a one-form on ${\mathcal M}_H$.) However, the
form $\omega_I$ is topologically non-trivial. It is given by the
imaginary part of (\ref{zork}), or
\begin{equation}\label{obox}\omega_I=\frac{1}{4\pi}\int_C{\mathrm {Tr}}\,\,\left(\delta
A\wedge \delta A-\delta\phi\wedge\delta\phi\right).\end{equation}
One way to prove that $\omega_I$ has a non-zero cohomology class
is to observe that, when restricted to the locus $\phi=0$, it
becomes the Kahler form of the compact Kahler manifold
${\mathcal M}\subset{\mathcal M}_H$ that parametrizes flat $G$-bundles on $C$. The
Kahler form of a compact Kahler manifold always has a non-trivial
cohomology class.
\def{\cmmib u}\,{{\cmmib u}\,}
\def{\cmmib v}\,{{\cmmib v}\,}
Now let us discuss what two-dimensional $A$-models we can make
using the symplectic forms $\omega_I$ and $\omega_K$ that do not
depend on a choice of complex structure on $C$. (We keep away
from $\omega_J$, since it does depend on the complex structure of
$C$.) Superficially, we can introduce two complex parameters
${\cmmib u}\,,{\cmmib v}\,$, since the complexified Kahler class $\widehat\omega$
used in defining the $A$-model may be any complex linear
combination
\begin{equation}\label{oplo}\widehat\omega={\cmmib u}\,
\omega_I+{\cmmib v}\,\omega_K.\end{equation} We must take the real parts
of ${\cmmib u}\,$ and ${\cmmib v}\,$ to be not both zero, because to define
an $A$-model, ${\mathrm{Re}}\,\widehat\omega$ must be a symplectic
form. However, in the absence of hypermultiplet bare masses,
because $\omega_K$ is exact, there is really only one complex
parameter that matters, namely ${\cmmib u}\,$. If
$\mathrm{Re}\,{\cmmib u}\,=0$, we must take
$\mathrm{Re}\,{\cmmib v}\,\not=0$, but its magnitude and even sign are
not relevant. (The sign of $\mathrm{Re}\,{\cmmib v}\,$ can be reversed
by a rotation of the Higgs field.) So really the only meaningful
parameter is ${\cmmib u}\,$. Arbitrary values of ${\cmmib u}\,\in{\Bbb C}$ make
sense, with ${\cmmib v}\,$ turned on if necessary (or if desired). There
is actually also a limit as ${\cmmib u}\,\to\infty$; this limit is the
$B$-model in complex structure $J$. The fact that the limit of
$A$-models for $\widehat\omega\to\infty$ exists and is a $B$-model is
most readily understood using generalized complex geometry and
will be reviewed in section \ref{rolegen}.
Both in interpreting geometric Langlands duality via gauge theory
\cite{KW} and in section \ref{compom} of the present paper, the
important special case of the $A$-model is the case ${\cmmib u}\,=0$.
This is naturally called the $A$-model of type $\omega_K$; as we
have explained, in an important situation, this model has no
Kahler parameter. In an extended version of geometric Langlands
duality, analyzed in section 11 of \cite{KW} and encountered in
section \ref{confblocks} of the present paper, one requires the
generic model with variable ${\cmmib u}\,$. This model is naturally
called the $A$-model of type $\omega_I$, and of course, it does
always have a Kahler parameter, namely ${\cmmib u}\,$.
We will make one last comment on branes of type $(A,B,A)$.
Although the conditions that characterize a brane of type
$(A,B,A)$ do not depend on a choice of complex structure on $C$, a
particular $(A,B,A)$-brane might be defined in a way that does
depend on that complex structure. Indeed, we have already
discussed a very important example. Given a complex structure on
$C$, ${\mathcal M}_H$ becomes hyper-Kahler, and in particular it acquires a
symplectic form $\omega_J$ that is Kahler with respect to $J$.
Unlike $\omega_I$ and $\omega_K$, $\omega_J$ does depend on the
metric of $C$ (as do $I=\omega_I^{-1}\omega_J$ and
$K=\omega_K^{-1}\omega_J$). So the canonical coisotropic
$(A,B,A)$ brane does depend on the complex structure of $C$,
though it is a brane in a family of topological field theories
that do not depend on this metric.
\subsection{A Little Differential Geometry}\label{little}
On a hyper-Kahler manifold $X$, we have described some branes
with multiple supersymmetric properties -- the brane ${\mathcal B}_*$ is of
type $(B,B,B)$, and its cousin ${\mathcal B}_{\mathrm{cc}}$ is of type $(A,B,A)$. These
are not the only examples. Another simple example of a brane of
type $(B,B,B)$ is a brane supported at a point in $X$ -- since a
point is a complex submanifold in any complex structure. The most
obvious branes of type $(A,B,A)$ are Lagrangian branes of this
type. Such a brane is supported on a middle-dimensional
submanifold $L\subset X$ that is holomorphic in complex structure
$J$, and is Lagrangian for the holomorphic symplectic form
$\Omega_J=\omega_K+i\omega_I$. As $L$ is Lagrangian for both
$\omega_K$ and $\omega_I$, a brane supported on $L$ with vanishing
Chan-Paton curvature is an $A$-brane of these types; as $L$ is
holomorphic in complex structure $J$, such a brane is of type
$(A,B,A)$.
The structures that we have described are redundant, in the
following sense. A brane that is (for example) a $B$-brane of
type $I$ and a $B$-brane of type $J$ is automatically a $B$-brane
of type $K$ (and more generally, a $B$-brane in any complex
structure $aI+bJ+cK$, $a^2+b^2+c^2=1$), since the conserved
supercharges of these three $B$-models are linearly dependent.
This linear dependence will become very clear in section
\ref{compom}, but here we will briefly explain the redundancy
among the different supersymmetric structures from the point of
view of differential geometry.
In a hyper-Kahler manifold $X$, consider a brane ${\mathcal B}$ with support
$L\subset X$. One condition for ${\mathcal B}$ to be a $B$-brane for
complex structures $I$ and $J$ is that $L$ must be holomorphic in
those complex structures. If so, then $L$ is also holomorphic in
complex structure $K=IJ$. Indeed, if the tangent space to $L$ is
invariant under the endomorphisms of the tangent bundle to $X$
corresponding to $I$ and $J$, it is certainly invariant under
$K=IJ$. The other condition for ${\mathcal B}$ to be a $B$-brane for
complex structures $I$ and $J$ is that the Chan-Paton curvature
$F$ is of type $(1,1)$ with respect to both $I$ and $J$;
equivalently, $I^tFI=J^tFJ=F$. Clearly this implies that
$K^tFK=F$, completing the argument that ${\mathcal B}$ is a $B$-brane of
type $K$ if it is one of types $I$ and $J$.
For an analogous argument for $A$-branes of type $(A,B,A)$, we
will consider just the case of a Lagrangian brane ${\mathcal B}$ supported
on a middle-dimensional submanifold $L$. We will show that if
${\mathcal B}$ is an $A$-brane for both $\omega_I$ and $\omega_K$, then it
is a $B$-brane in complex structure $J$. (We leave it to the
reader to show that if ${\mathcal B}$ is an $A$-brane for $\omega_I$ and a
$B$-brane for $J$, then it is an $A$-brane for $\omega_K$.) Let
$TL$ be the tangent bundle to $L$, and let $N^*L$ be the subspace
of $T^*X|_L$ (the restriction to $L$ of the cotangent bundle of
$X$) consisting of cotangent vectors that annihilate $TL$. The
fact that $L$ is Lagrangian for both $\omega_I$ and $\omega_K$
means that $\omega_I$ establishes an isomorphism from $TL$ to
$N^*L$, and $\omega_K^{-1}$ is an isomorphism from $N^*L$ to $TL$.
So $J=\omega_K^{-1}\omega_I$ is an isomorphism from $TL$ to
itself, and thus $L$ is holomorphic in complex structure $J$ and
${\mathcal B}$ is a $B$-brane.
In particular (though we have only shown this for Lagrangian
branes) there is no such thing as a brane of type $(A,A,A)$ -- if
${\mathcal B}$ is an $A$-brane of type $I$ and $K$, then it is a $B$-brane
of type $J$. To illuminate the last statement further, and for
some further applications, we will give an overview of the
possible half-BPS supersymmetry conditions for a brane on a
hyper-Kahler manifold $X$.
\subsubsection{General Half-BPS Condition}\label{genhalf}
\def{\mathcal J}{{\mathcal J}}
On $X$, there is a family of complex structures parametrized by
$\mathbb{CP}^1$; a general element of this family is a complex
structure ${\mathcal J}=aI+bJ+cK$, with $a,b,c$ real and $a^2+b^2+c^2=1$.
Twisted topological field theories in two dimensions are
conveniently constructed by twisting a theory with $(2,2)$
supersymmetry. In general, a sigma-model with target $X$ and
$(2,2)$ supersymmetry is constructed \cite{Rocek} in terms of a
pair of integrable complex structures ${\mathcal J}_+$ and ${\mathcal J}_-$, which
govern right- and left-moving excitations, respectively. They
obey a certain compatibility condition which also involves the
metric and the curvature $H$ of the $B$-field. Generalized
complex geometry \cite{Hitchin2} leads to the most elegant
interpretation of the compatibility condition \cite{Gualtieri}. We
use this viewpoint below.
If $X$ is a hyper-Kahler manifold, ${\mathcal J}_+$ and ${\mathcal J}_-$ can be chosen
to correspond to arbitrary points in $\mathbb{CP}^1$; the
compatibility condition is always obeyed, with $H=0$. Hence, the
sigma-model with hyper-Kahler target space has a twisted version
corresponding to an arbitrary pair $({\mathcal J}_+,{\mathcal J}_-)\in
\mathbb{CP}^1_+\times \mathbb{CP}^1_-$, that is, in the product of two
copies of $\mathbb{CP}^1$. A $B$-model corresponds to the case that
${\mathcal J}_+={\mathcal J}_-$, and an $A$-model corresponds to the case that
${\mathcal J}_+=-{\mathcal J}_-$.
A supersymmetric boundary condition preserves the supersymmetries
associated to certain pairs $({\mathcal J}_+,{\mathcal J}_-)$, but of course, not all
possible pairs. In general, a half-BPS boundary condition
preserves the supersymmetries associated with pairs of the form
${\mathcal J}_-=h{\mathcal J}_+$, where $h\in
SO(3)$ is a rigid rotation of $\mathbb{CP}^1\cong S^2$ that gives a
holomorphic map from $\mathbb{CP}^1_+$ to $\mathbb{CP}^1_-$. For
example, if $h=1$, we have ${\mathcal J}_+={\mathcal J}_-$ for all ${\mathcal J}_+$. This is the
condition for a brane of type $(B,B,B)$. Since the antipodal map
on $\mathbb{CP}^1$ is not as $SO(3)$ rotation, it is not possible to
have ${\mathcal J}_-=-{\mathcal J}_+$ for all ${\mathcal J}_+$, and hence there is no such thing
as a brane of type $(A,A,A)$.
\def\varrho{\varrho}
Actually, any $h\in SO(3)$ leaves fixed some axis in
$\mathbb{CP}^1\cong S^2$, so there is always some choice of ${\mathcal J}_+$
for which $h{\mathcal J}_+={\mathcal J}_+$. Hence any half-BPS brane ${\mathcal B}$ is a
$B$-brane in some complex structure ${\mathcal J}=aI+bJ+cK$, with
$a^2+b^2+c^2=1$. It is not true that there is always some ${\mathcal J}_+$
with $h{\mathcal J}_+=-{\mathcal J}_+$. Such a ${\mathcal J}_+$ exists if and only if $h$ is a
$\pi$ rotation around the appropriate axis. If so, then regarding
$h$ as a linear transformation of a copy of ${\mathbb R}^3$ in which
$S^2\cong\mathbb{CP}^1$ is embedded, $h$ has two eigenvalues $-1$,
and so a brane associated with such a $h$ is an $A$-brane in two
different ways. After a suitable rotation of the coordinate axes
(so that $h$ is a rotation around the $J$ axis), such a brane is
of type $(A,B,A)$.
\subsubsection{Role Of Generalized Complex
Geometry}\label{rolegen}
But what happens if $h$ is a rotation by an angle other than
$\pi$? In this case, although ${\mathcal B}$ is a $B$-brane in one complex
structure, its other supersymmetric properties appear unfamiliar.
To allow for the case of an arbitrary $h$, we consider the general
case of a brane ${\mathcal B}$ that conserves a topological supercharge $Q$
associated to a pair of independent complex structures ${\mathcal J}_+$ and
${\mathcal J}_-$ for the right-moving and left-moving modes. It turns out
\cite{Kap} that on a hyper-Kahler manifold, the topological field
theory associated to a pair $({\mathcal J}_+,{\mathcal J}_-)$ with ${\mathcal J}_+\not={\mathcal J}_-$ can
always be reduced to an $A$-model, even if ${\mathcal J}_+\not={\mathcal J}_-$. The
reduction is made using the language of generalized complex
geometry \cite{Hitchin2,Gualtieri}. (The requisite formulas are
summarized in section 5.2 of \cite{KW}, where they are applied to
geometric Langlands.)
\def{\mathcal M}{{\mathcal M}}
Rather than defining a topological twist by a pair of complex
structures with a metric and $B$-field obeying certain conditions,
a useful point of view is that such a twist can be determined by
the choice of a generalized complex structure ${\mathcal I}$. A generalized
complex structure is a linear transformation ${\mathcal I}$ of $TX\oplus
T^*X$ (the direct sum of the tangent and cotangent bundles of $X$)
that obeys ${\mathcal I}^2=-1$ as well as a certain integrability condition.
A $B$-model associated to a complex structure $J$ (whose transpose
we denote as $J^t$) corresponds to the case that
\begin{equation}\label{mlef}{\mathcal I}_J=\begin{pmatrix} J & 0 \cr 0 &
-J^t\end{pmatrix}.\end{equation}
The $A$-model with a symplectic structure $\omega$ and zero $B$-field
corresponds to the case that
\begin{equation}\label{nlef}{\mathcal I}_\omega=\begin{pmatrix}0 & -\omega^{-1}\cr \omega& 0\end{pmatrix}.\end{equation}
In general, to turn on a $B$-field, pick a closed two-form $B_0$ and set
\begin{equation}\label{plef}{\mathcal M}(B_0)=\begin{pmatrix} 1 & 0 \cr B_0 & 1\end{pmatrix}.\end{equation}
The transformation
\begin{equation}\label{qlef}{\mathcal I}\to {\mathcal M}(B_0){\mathcal I}{\mathcal M}(B_0)^{-1}\end{equation}
is known as a $B$-field transform. It
preserves the condition ${\mathcal I}^2=-1$ and the integrability condition obeyed by ${\mathcal I}$, and has the effect
of shifting the $B$-field by $B_0$. In particular, the generalization of ${\mathcal I}_\omega$ to include a $B$-field
is
\begin{equation}\label{mulef}{\mathcal I}_{\omega,B}={\mathcal M}(B)\begin{pmatrix}0 & -\omega^{-1}\cr \omega& 0\end{pmatrix}
{\mathcal M}(B)^{-1}.\end{equation}
Now return to the case of a hyper-Kahler manifold $X$ with a pair
of complex structures ${\mathcal J}_\pm$ for right-movers and left-movers.
Denoting as $g$ the hyper-Kahler metric of $X$, let
$\omega_\pm=g{\mathcal J}_\pm$ be the Kahler forms associated to the complex
structures ${\mathcal J}_\pm$. The generalized complex structure associated
to this data is, according to eqn. 6.3 of \cite{Gualtieri},
\begin{equation}\label{nulef}{\mathcal J}=\frac{1}{2}\begin{pmatrix}{\mathcal J}_++{\mathcal J}_-
& -(\omega_+^{-1}-\omega_-^{-1})\cr
\omega_+-\omega_-&-(J_+^t+J_-^t)\end{pmatrix}.\end{equation}
(There is also a second generalized complex structure that we do
not need here; it is obtained by reversing the sign of ${\mathcal J}_-$ and
$\omega_-$.) As long as ${\mathcal J}_+\not={\mathcal J}_-$, this takes the form of
eqn. (\ref{mulef}) with
\begin{align}\label{golf}\omega^{-1}&=\frac{1}{2}\left(\omega_+^{-1}-\omega_-^{-1}\right)
\cr \omega^{-1}B&=\frac{1}{2}({\mathcal J}_++{\mathcal J}_-).\end{align}
Hence, the model is equivalent to an $A$-model.
For our application, an important special case is that ${\mathcal J}_+$ is
very close to ${\mathcal J}_-$ -- so the model is almost a $B$-model. If
${\mathcal J}_+-{\mathcal J}_-$ is of order $\varepsilon$, where $\varepsilon$ is a
small parameter, then according to (\ref{golf}), $\omega$ and $B$
and therefore the complexified Kahler form $\widehat\omega=\omega+iB$
are of order $\varepsilon^{-1}$. This seems a little puzzling
because one expects the effects of rotating ${\mathcal J}_+$ slightly away
from ${\mathcal J}_-$ to be small. However, as we have reviewed in section
\ref{coiso}, noncommutative effects in the $A$-model are of order
$\widehat\omega^{-1}$, which in the present context means that these
effects are of order $\varepsilon$. In the limit that ${\mathcal J}_+$
approaches ${\mathcal J}_-$, the noncommutative effects in the $A$-model
vanish and the $A$-model becomes an ordinary commutative
$B$-model.
Since $\widehat\omega$ diverges as $\varepsilon\to 0$, one might not
expect the $A$-model with complexified symplectic form
$\widehat\omega$ to have a limit as $\varepsilon\to 0$. But in fact
this limit exists and is simply the $B$-model of type ${\mathcal J}_+$ or
equivalently ${\mathcal J}_-$.
\subsubsection{``Rotation'' Group}\label{rotgroup}
\def{\Gamma}{{\Gamma}}
\def{\Bbb{CP}}{{\mathbb{CP}}}
Two-dimensional topological field theories of the class considered
here are labeled by the pair $({\mathcal J}_+,{\mathcal J}_-)$, which parametrize what
we may call ${\Bbb{CP}}^1_+\times {\Bbb{CP}}^1_-$, with one copy of ${\Bbb{CP}}^1$ for
${\mathcal J}_+$ and one for ${\mathcal J}_-$. It is natural to introduce a group
${\Gamma}=SU(2)_+\times SU(2)_-$ that rotates ${\Bbb{CP}}^1_+\times {\Bbb{CP}}^1_-$,
with one factor of $SU(2)$ for each factor of ${\Bbb{CP}}^1$. The group
that acts faithfully on ${\Bbb{CP}}^1_+\times {\Bbb{CP}}^1_-$ is actually
$SO(3)_+\times SO(3)_-$, where $SO(3)_\pm = SU(2)_\pm/{\Bbb Z}_2$. ${\Gamma}$
is a double cover of $SO(4)=(SU(2)_+\times SU(2)_-)/{\Bbb Z}_2$.
Consider a half-BPS brane characterized by a condition
${\mathcal J}_-=h{\mathcal J}_+$, $h\in SU(2)$. Obviously, if we transform
$({\mathcal J}_+,{\mathcal J}_-)$ to $(g_+{\mathcal J}_+,g_-{\mathcal J}_-)$, then $h$ is transformed to
$h'=g_-^{-1}hg_+$.
For example, suppose that $h=1$, corresponding to a brane of type
$(B,B,B)$. Then $h'=g_-^{-1}g_+$. If (as will occur in our
application), $g_+$ is a rotation around some axis by an angle
$\vartheta$, and $g_-=g_+^{-1}$, then $h'$ is a rotation around
the given axis by an angle $2\vartheta$.
\section{Compactification And $\Omega$-Deformation}\label{compom}
Finally, we are prepared to consider our first application: the
relation of the $\Omega$-deformation to quantization.
The object of study in \cite{NS2} was a four-dimensional gauge
theory with ${\mathcal N}=2$ supersymmetry, ``compactified'' to two
dimensions on ${\mathbb R}^2_\varepsilon$. Here ${\mathbb R}^2_\varepsilon$ is simply
${\mathbb R}^2$, endowed with a $U(1)$ rotation symmetry that leaves the
origin fixed; the gauge theory on ${\mathbb R}^2_\varepsilon$ is deformed via
the $\Omega$-deformation \cite{Nek} with parameter $\varepsilon$.
The precise metric on ${\mathbb R}^2$ is not essential, as long as it is
$U(1)$-invariant. We will find it helpful to place on ${\mathbb R}^2$ a
``cigar-like'' metric
\begin{equation}\label{cigar}{\mathrm d} s^2={\mathrm d} r^2+
f(r)\,{\mathrm d}\theta^2,~0\leq r<\infty,~~0\leq\theta\leq
2\pi,\end{equation} with $f(r)\sim r^2$ for $r\to 0$ and $f(r)\to
\rho^2$ for $r\to\infty$. Thus $\rho$ is the asymptotic radius of
the circle parametrized by $\theta$. We can assume that $f(r)$ is
identically equal to $\rho$ for sufficiently large $r$ (say $r\geq
r_0$). We write $D$ for ${\mathbb R}^2$ endowed with this kind of metric.
We also write $D_R$ for $D$ restricted to $r\leq R$, where we
choose $R$ so that $R>>\rho,r_0$.
We will compactify to two dimensions on $D_R$,
with an $\Omega$-deformation and a
suitable supersymmetric boundary condition at $r=R$. However,
first we will need to understand what happens in the absence of
the $\Omega$-deformation.
To make contact with the explanation of integrability in section \ref{linko}, we take
the two-manifold to which we compactify on $D_R$ to be ${\mathbb R}\times S^1$.
So overall, we will be doing gauge theory on $M={\mathbb R}\times S^1\times D_R$.
Since we take the cutoff $R$ very large, $D_R$ looks macroscopically like $I\times
\widetilde S^1$, where $\widetilde S^1$ is a second circle, parametrized by $\theta$, and the interval
$ I$ is parametrized by $r$, $0\leq
r\leq R$. Macroscopically, $M$ is a two-torus fibration, that is
an $S^1\times \widetilde S^1$ fibration, over ${\mathbb R}\times I$. We should
be able to reduce to an effective description in a sigma-model on
${\mathbb R}\times I$.
\def{\cmmib B}{{\cmmib B}}
\def{\mathcal M}{{\mathcal M}}
The appropriate sigma-model is obtained by compactification of our
four-dimensional gauge theory to two dimensions on a two-torus
$T^2$. For orientation, we consider the generalized quiver
theories that are obtained \cite{DG} by compactifying the
six-dimensional $(0,2)$ theory on a Riemann surface $C$, perhaps
with surface operators supported at marked points on $C$. In this
case, our ${\mathcal N}=2$ gauge theory on $M={\mathbb R}\times S^1\times D_R$ will
reduce at long distances to the sigma-model on $\Sigma={\mathbb R}\times I$
with target ${\mathcal M}_H$, the moduli space of Higgs bundles on $C$.
To complete this description, we need to specify two branes,
supplying boundary conditions at the two ends of $I$. The brane at
$r=0$ will in some sense arise purely from geometry, as $r=0$ is
not really a boundary point in the more microscopic description on
$M$. So one of our questions will be to identify the brane that
is generated by geometry. The second boundary condition in the
two-dimensional description, the one at $r=R$, will descend from a
choice of a boundary condition in the four-dimensional gauge
theory.
To account for the results of \cite{NS2}, we want quantization of
the sigma-model on $\Sigma={\mathbb R}\times I$ to give quantization in the
ordinary sense of a middle-dimensional real subspace of ${\mathcal M}_H$.
From section \ref{coiso}, we know how this might happen: the
effective model on $\Sigma$ should be a two-dimensional $A$-model;
one brane should be a canonical coisotropic brane ${\mathcal B}_{\mathrm{cc}}$, with
support all of ${\mathcal M}_H$, while the other should be an ordinary
Lagrangian brane ${\mathcal B}_L$, with support a Lagrangian submanifold
$L\subset{\mathcal M}_H$. It will turn out that the brane that arises from
geometry will be ${\mathcal B}_{\mathrm{cc}}$, while the Lagrangian brane ${\mathcal B}_L$ will
depend on a choice of boundary condition at $r=R$.
In section \ref{undeformed}, we study compactification on $D_R$ in
the absence of the $\Omega$-deformation. We identify the brane
that arises at $r=0$ in the effective two-dimensional description.
The support of this brane is all of ${\mathcal M}_H$; however, it is not a
coisotropic brane, but the more elementary brane ${\mathcal B}_*$ of type
$(B,B,B)$ described at the end of section \ref{hyperkahler}. In
view of \cite{NS2}, the way to remedy this must be to incorporate
the $\Omega$-deformation. In section \ref{rethink}, we reformulate
the $\Omega$-deformation in a way suitable for our purposes. In
section \ref{deformed}, we consider the $\Omega$-deformed theory
on $M={\mathbb R}\times I\times D_R$ and explain why the
$\Omega$-deformation has the desired effects. In section
\ref{farend}, we describe boundary conditions at the far end of
$D_R$.
\subsection{The Undeformed Case}\label{undeformed}
Two different $2+2$-dimensional splits of $M={\mathbb R}\times S^1\times D_R$
will be important in this paper. The first is the obvious decomposition of $M$
as the product of two two-manifolds ${\mathbb R}\times S^1$ and $D_R$. The second
involves using the fact that $D_R$ is asymptotic to $I\times \widetilde S^1$ and viewing $M$ as an $S^1\times \widetilde S^1$ fibration over ${\mathbb R}\times I$.
Unfortunately, it is difficult to find a notation that is well-adapted to both decompositions.
What we will do is simply to number the coordinates as $0,1,2,3$ for ${\mathbb R}$, $S^1$,
$I$, and $\widetilde S^1$, respectively.
The bosonic part of the four-dimensional vector multiplet
comprises a gauge field $A_\mu$, $\mu=0,\dots,3$ and a complex
scalar $\phi$ in the adjoint representation. It is convenient to
adopt a six-dimensional notation in which $A_\mu$ and $\phi$
combine to a six-dimensional gauge field $A_I$, $I=0,\dots,5$
which is independent of the last two coordinates; from this point
of view, $\phi=(A_4-iA_5)/\sqrt 2$. This is useful because,
although there is not really an $SO(6)$ symmetry rotating the six
components of $A_I$, many key equations can conveniently be
written in $SO(6)$ notation. For example, the supersymmetry
generator $\eta$ is a spinor of definite chirality, so if we
introduce gamma matrices $\Gamma_I$, $I=0,\dots,6$, obeying (in
Euclidean signature) $\{\Gamma_I,\Gamma_J\}=2\delta_{IJ}$, then
\begin{equation}\label{turmok}\Gamma_0\Gamma_1\cdots\Gamma_5\eta
= i\eta.\end{equation} The fermions $\Psi$ of the vector multiplet
are similarly a Weyl spinor in the adjoint representation of the
gauge group, with the same chirality as $\eta$. Apart from being
chiral spinors of $SO(6)$, $\eta$ and $\Psi$ are also spinors of
the $SU(2)_R$ group of $R$-symmetries. We denote as $\sigma_i$,
$i=1,2,3$ the analogs for $SU(2)_R$ of the gamma matrices, obeying
$\sigma_i\sigma_j=\delta_{ij}+i\epsilon_{ijk}\sigma_k$. We will
use standard abbreviations such as $\Gamma_{IJ}=\Gamma_I\Gamma_J$,
$I\not=J$, and
$\sigma_{ij}=\sigma_i\sigma_j=i\epsilon_{ijk}\sigma_k$, $i\not=j$.
An example of the usefulness of the six-dimensional notation is
that the supersymmetry transformations for the vector multiplet
are simply written:
\begin{align}\label{urkok}\delta A_I& = i\overline\eta\Gamma_I\Psi \cr \delta\Psi & = \frac{1}{2}\Gamma^{IJ}F_{IJ}\eta.\end{align}
Now consider an ${\mathcal N}=2$ gauge theory on the four-manifold
$M={\mathbb R}\times S^1\times D_R$. Away from the tip of the cigar (that
is, the region near $r=0$), $D_R$ is equivalent to $I\times
\widetilde S^1$ and so $M$ reduces to the flat manifold ${\mathbb R}\times
S^1\times I\times \widetilde S^1$. On this flat manifold, there are
eight unbroken supersymmetries corresponding to all eight
components of $\eta$.
The curvature near the tip of the cigar inevitably breaks some of
the supersymmetries, in fact at least half of them. (Any set of
at least five supersymmetries would include one whose square would
generate in the asymptotic region a translation along the first
factor of $D_R\sim I\times \widetilde S^1$, but such a translation
cannot be extended to a symmetry of $D_R$.) There is a standard
way \cite{WD} to make a topological twist so that half of the asymptotic
supersymmetries are preserved in the exact $D_R$ geometry. The
supersymmetries that are preserved are the ones that are invariant
under a rotation of the tangent space of $I\times \widetilde S^1$
together with an $SU(2)_R$ rotation. The rotation of the tangent
space of $I\times \widetilde S^1$ is generated by $\Gamma_{23}$, and
up to conjugation in $SU(2)_R$, we can assume that the $SU(2)_R$
rotation in question is generated by $\sigma_{23}$. So with the
standard topological twist, the four supersymmetries that are
preserved are the ones that can be characterized, in the
asymptotic region of $D_R$, by
\begin{equation}\label{zomely}\left(\Gamma_{23}+\sigma_{23}\right)\eta=0.\end{equation}
Let us look at this from the point of view of toroidal
compactification, on $S^1\times \widetilde S^1$, to $\Sigma={\mathbb R}\times
I$. The tip of the cigar at $r=0$ gives a boundary condition at
one end of $I$. This boundary condition preserves half of the
supersymmetry. In other words, in the effective two-dimensional
sigma-model, the tip, with the standard topological twist,
determines a half-BPS brane ${\mathcal B}_*$. We would like to interpret in
two-dimensional terms the unbroken supersymmetry and the brane
that carries this symmetry. A convenient way to do this is to
understand what topological properties this brane possesses. What
structures of twisted topological field theory on $\Sigma$ are
preserved by this brane?
Any topological field theory structure on $\Sigma$ is associated
with a supersymmetry generator $\eta$ that is invariant under
a rotation of the tangent space to $\Sigma$ together with an
$SU(2)_R$ transformation. The rotation of the tangent space is
generated by $\Gamma_{02}$, and as this anticommutes with the
matrix $\Gamma_{23}$ that appears on the left of eqn.
(\ref{zomely}), we must pick an $SU(2)_R$ generator that
anticommutes with $\sigma_{23}$ or we will reach a contradiction.
With no essential loss of generality, we can look for a
supersymmetry generator that obeys
\begin{equation}\label{pomely}\left(\Gamma_{02}+\sigma_{31}\right)\eta=0.\end{equation}
The two equations (\ref{zomely}) and (\ref{pomely}) characterize a
two-dimensional space of $\eta$'s. To determine a particular
topological field theory structure on $\Sigma$, we need one more
condition restricting to a one-dimensional space of $\eta$'s.
Any condition will do, so there is a $\mathbb{CP}^1$ family of
topological field theories on $\Sigma$ that are all compatible
with the same brane ${\mathcal B}_*$.
For one convenient choice, we supplement (\ref{zomely}) and
(\ref{pomely}) with the additional condition
\begin{equation}\label{tomely}\left(\Gamma_{01}+\sigma_{23}\right)\eta=0.\end{equation}
Although presented here in a non-invariant way, these three
conditions combine to something that can be described invariantly.
By commuting the operators appearing on the left hand sides of the
three equations, one learns that a spinor obeying the three
equations actually obeys
\begin{align} \label{fomely}
\left(\Gamma_{ij}+\sigma_{ij}\right)\eta& = 0\cr
\left(\Gamma_{0i}+\sigma_{i+1,i-1}\right)\eta& = 0,
\end{align}
for $i,j=1,2,3$. By subtracting (\ref{tomely}) from (\ref{zomely}), we find that
$\left(\Gamma_{01}-\Gamma_{23}\right)\eta=0$, which together with (\ref{turmok})
implies that
\begin{equation}\label{gelmok}\Gamma_{0123}\eta=-\eta,~~\Gamma_{45}\eta=-i\eta.\end{equation}
The conditions (\ref{fomely}) are the standard conditions
that characterize the supersymmetry generator of a twisted
four-dimensional topological field theory \cite{WD} that (in the
case of $SU(2)$ gauge theory without hypermultiplets) is related
to Donaldson theory. They can be characterized in group-theoretic terms.
Let $SO(4)$ be the group of rotations of the
tangent space to $M$; denote its double cover as $SU(2)_l\times
SU(2)_r$. Then the above conditions mean that $\eta$ is invariant under
$SU(2)_l\times SU(2)'_r$, where $SU(2)'_r$ is a diagonal subgroup
of $SU(2)_r\times SU(2)_R$.
We can get two more useful choices of supersymmetry parameter by
observing that $\Gamma_1$, $\Gamma_4$, and $\Gamma_5$ all commute
with the operators on the left hand sides of our first two
conditions (\ref{zomely}) and (\ref{pomely}). So these conditions
commute with a group $SO(3)_{145}$ that rotates those three gamma
matrices. Making an $SO(3)_{145}$ transformation that rotates
$\Gamma_1$ to $\Gamma_4$ or $\Gamma_5$, we replace (\ref{tomely})
by
\begin{equation}\label{domely}\left(\Gamma_{04}+\sigma_{23}\right)\eta=0\end{equation}
or
\begin{equation}\label{bomely}\left(\Gamma_{05}+\sigma_{23}\right)\eta=0.\end{equation}
\subsubsection{Support Of The Brane}\label{otto}
Before trying to interpret the supersymmetries in two dimensional
terms, let us first determine the support of the brane ${\mathcal B}_*$.
In four dimensions, a single vector multiplet contains a complex
scalar field $\phi$ or equivalently a pair of real scalars. When
we compactify to three dimensions on a circle $S^1$, two more
scalars come from the gauge field $A$ -- one is the holonomy of
$A$ around $S^1$, and the second is the dual photon. All four
scalars combine to a three-dimensional hypermultiplet. The key
point here is that to arrive at this hypermultiplet, a duality
transformation was needed, converting the photon to a scalar.
The geometry of the hypermultiplets that arise in compactification
can be described as follows. Let $\cmmib B$ be the Coulomb branch
of vacua of the four-dimensional gauge theory. It parametrizes a
family of abelian varieties. We denote the total space of this
family as ${\mathcal M}_H$ because in a large class of examples \cite{DG},
this total space is a moduli space of Higgs bundles on a Riemann
surface $C$. After compactification on a circle and dualization
of the photons (one in each vector multiplet), ${\mathcal M}_H$ becomes
endowed with a hyper-Kahler metric, and one gets \cite{SW2} a low
energy description by a sigma-model of maps from three-dimensional
spacetime to ${\mathcal M}_H$.
Compactification to two dimensions on $S^1\times \widetilde S^1$ is a
little different. In this case, a four-dimensional gauge field
leads to two scalars -- its holonomies around the two circles --
without dualization of any kind. This actually gives a
description by linear multiplets rather than hypermultiplets
\cite{Rocek}; this description is inconvenient as there is not a
powerful theory of nonlinear models built from linear multiplets.
A $T$-duality transformation for the scalar fields that arise from
holonomies around one circle or the other is useful because it
leads to a description by hypermultiplets, and here hyper-Kahler
geometry is an effective tool for studying nonlinear models. (For
more on this, see the end of this subsection as well as section
\ref{mixed}.)
From a two-dimensional point of view, in compactification on
$S^1\times \widetilde S^1$, there is no natural choice of which of the
two sets of scalars should be $T$-dualized. A description in
which we $T$-dualize one set of scalars differs from a description
in which we $T$-dualize the other set of scalars by a combined
$T$-duality on both sets of scalars. The combined operation is a
$T$-duality on all the scalars that come from gauge fields, so it
can be described simply: it is the $T$-duality on the fibers of
the fibration ${\mathcal M}_H\to \cmmib B$. This particular instance of
$T$-duality is related to $S$-duality in another description of
the same models \cite{HM,BJV}, and is the basic geometric
Langlands duality \cite{KW}. This $T$-duality transforms ${\mathcal M}_H$
to an analogous moduli space of Higgs bundles for the Langlands
dual gauge group.
Now let us specialize to our problem with $M={\mathbb R}\times S^1\times
D_R\sim{\mathbb R}\times S^1\times I\times \widetilde S^1$. The symmetry
between $S^1$ and $\widetilde S^1$ is broken by the fact that $\widetilde
S^1$, and not $S^1$, is capped off at the tip of the cigar. It
turns out that to explain the results of \cite{NS2}, it is better
to $T$-dualize the holonomies of the gauge field around $\widetilde
S^1$. (In section \ref{confblocks}, we will explore another
problem in which the two circles enter symmetrically.)
Recalling that we have labeled the four dimensions of $M\sim
{\mathbb R}\times S^1\times I\times \widetilde S^1$ consecutively as 0123, we
write simply $A_1$ or $A_3$ for scalars arising from the holonomy
of a gauge field $A$ around $S^1$ or $\widetilde S^1$, respectively.
The boundary conditions on $\phi$, $A_1$, and $A_3$ at the
tip of the cigar are uniquely determined, since in
four-dimensional terms there is no boundary at all. There is no
reason for $\phi$ or $A_1$ to vanish at the tip of the cigar, so
in the two-dimensional description on $\Sigma={\mathbb R}\times I$, they
obey Neumann boundary conditions. On the other hand, $A_3$ must
vanish because it is the holonomy of the gauge field around a
circle that shrinks to a point at the tip. So $A_3$ obeys
Dirichlet boundary conditions.
However, to get a description by a two-dimensional sigma-model
with target ${\mathcal M}_H$, we are supposed to $T$-dualize $A_3$,
replacing it by another scalar that we will call $\varrho$. For
future reference,\footnote{\label{firstone} Here we assume the
circles $S^1$ and $\widetilde S^1$ are orthogonal; otherwise $A_1$
enters the formulas. See the appendix for a much more complete
treatment.} we write the equations describing this $T$-duality:
\begin{align}\label{confo}\partial_0 \varrho &=- i\partial_2 A_3\cr
\partial_2\varrho & = i\partial_0 A_3.\end{align}
After $T$-duality, $\varrho$ obeys Neumann boundary conditions at
the tip of the cigar. In fact, at this stage all scalars
$\phi,A_1,$ and $\varrho$ obey Neumann boundary conditions at the
tip. So the tip of the cigar corresponds in two-dimensional terms
to a brane ${\mathcal B}_*$ whose support is all of ${\mathcal M}_H$.
The topological twist that was used to preserve supersymmetry on
${\mathbb R}\times S^1\times D_R$ does not generate at the tip of $D_R$ any
couplings that look obviously like Chan-Paton couplings. So it is
natural to think that the Chan-Paton bundle of ${\mathcal B}_*$ may be
trivial. If so, as the support of ${\mathcal B}_*$ (being all of ${\mathcal M}_H$) is
holomorphic in every complex structure on ${\mathcal M}_H$, ${\mathcal B}_*$ will be a
brane of type $(B,B,B)$ -- a $B$-brane for every complex structure
that makes up the hyper-Kahler structure of ${\mathcal M}_H$. We will show
in section \ref{twod} that this is the case.
We will add a word on the more naive description by scalars $\phi,
A_1$, and $A_3$ without any $T$-duality. In this description,
precisely one scalar in each hypermultiplet (namely $A_3$) obeys
Dirichlet boundary conditions. Since the usual supersymmetric
branes have Dirichlet boundary conditions for an even number of
scalars in each multiplet, this is another indication that a
simple description requires $T$-duality for either $A_1$ or $A_3$.
(Without such a $T$-duality, a four-dimensional vector multiplet
reduces in two dimensions to a linear multiplet rather than a
hypermultiplet.)
\subsubsection{Two-Dimensional Interpretation Of The
Supersymmetries}\label{twod}
Now we want to determine the two-dimensional interpretation of the
supercharges whose generator $\eta$ obeys (\ref{zomely}) and
(\ref{pomely}) plus one of the three supplementary conditions (\ref{tomely}),
(\ref{domely}), or (\ref{bomely}).
We first consider the case of a spinor $\eta$ obeying
(\ref{tomely}). We claim that $\eta$ generates the topological
symmetry of the $B$-model in complex structure $I$ -- the complex
structure in which the Hitchin fibration is holomorphic.
The basic functions on ${\mathcal M}_H$ that are holomorphic in complex
structure $I$ are $\phi$ and $A_1+i\varrho$. So, if $Q$ is the
supersymmetry generated by $\eta$, we must show that
$[Q,\phi]=[Q,A_1+i\varrho]=0$.
It is straightforward to show that $[Q,\phi]=0$; this is actually
a standard fact in the context of applications to Donaldson theory
\cite{WD}. As $\phi=(A_4-iA_5)/\sqrt 2$ and $\delta
A_I=i\overline\eta\Gamma_I \Psi$, what we need to show is that
$\overline\eta(\Gamma_4-i\Gamma_5)\Psi=0$, which will follow if
$(\Gamma_4-i\Gamma_5)\eta=0$. This is equivalent to
$\Gamma_{45}\eta =-i\eta$, which was deduced in (\ref{gelmok}).
The other condition $[Q,A_1+i\varrho]=0$ is more subtle, because
$\varrho$ is defined via a $T$-duality that only makes sense
after reduction to two dimensions. So in analyzing this condition,
we work in the effective two-dimensional theory. Thus, we discard
terms involving derivatives in the 1 or 3 directions, and for
example that means that $F_{01}$ reduces to $\partial_0 A_1$. The
zero mode of a scalar field such as $\varrho$ that is defined via
$T$-duality is subtle to understand. However, there are
straightforward formulas (\ref{confo}) for the derivatives of
$\varrho$. So we will content ourselves with showing the
vanishing of the derivatives along ${\mathbb R}\times I$ of
$[Q,A_1+i\varrho]$ in the effective two-dimensional theory. For
example, the derivative in the 0 direction is
$[Q,\partial_0A_1+i\partial_0\varrho]=[Q,\partial_0
A_1+\partial_2A_3] =[Q,F_{01}+F_{23}]$. This vanishes; indeed,
the combination $F_{01}+F_{23}$ is self-dual and therefore is
$Q$-exact and in particular $Q$-closed in the topological field
theory related to Donaldson theory. The derivative of
$[Q,A_1+i\varrho]$ in the 2 direction vanishes similarly.
What we have learned then is that if we select a spinor $\eta$
using the supplementary condition (\ref{tomely}), we get the
topological supercharge of the $B$-model of ${\mathcal M}_H$ in complex
structure $I$. This is the complex structure in which the Hitchin
fibration ${\mathcal M}_H\to \cmmib B$ is holomorphic, and the scalar fields
that are functions on $\cmmib B$ are likewise holomorphic.
The other conditions that we want to analyze, namely
(\ref{domely}) and (\ref{bomely}), can be formally obtained from
(\ref{tomely}), which we have just analyzed, by an $SO(3)_{145}$
transformation that exchanges $A_1$ with $A_4$ or $A_5$. This
fact can be used to interpret the supersymmetries in the low
energy theory without any computation. From each vector
multiplet, we get four scalars in the effective two-dimensional
sigma-model. Schematically we call them $A_1$, $A_4$, $A_5$, and
$\varrho$. Each set of four scalars forms a hypermultiplet whose
tangent space admits an action of the quaternion units $I,J,$ and
$K$. A formal $SO(3)_{145}$ rotation that exchanges $A_1$ with
$A_4$ or $A_5$, so as to map (\ref{tomely}) to (\ref{domely}) or
(\ref{bomely}), maps $I$ to $J$ or $K$. So while the auxiliary
condition (\ref{tomely}) determines $\eta$ to be the generator of
the $B$-model in complex structure $I$, (\ref{domely}) or
(\ref{bomely}) similarly determines $\eta$ to be the generator of
the $B$-model in complex structure $J$ or $K$, respectively.
At this level of generality, it is in part a convention which of the complex structures on ${\mathcal M}_H$
is called $J$ rather than $K$. In a large class of models \cite{DG} in which ${\mathcal M}_H$ actually is a moduli
space of Higgs bundles on a Riemann surface $C$, we can fix the definition of $A_4$ and $A_5$
and the conventions in the Higgs bundle equations so that
$J$ is the complex structure in which ${\mathcal M}_H$ parametrizes flat bundles on $C$ with complex
structure group. $K=IJ$ is then distinct from $J$ but equivalent to it by a $U(1)_R$ rotation, provided hypermultiplet
bare masses (which violate $U(1)_R$) are absent.
The notation just described is in accord with that
of \cite{Hitchin}.
\subsubsection{A Mixed $AB$-Model}\label{mixed}
Now we can be more precise about what would happen if we describe
the two-dimensional effective field theory with the naive set of
scalars $\phi$, $A_1$, and $A_3$, without any $T$-duality.
For example, let us consider the supercharge $Q$ of Donaldson
theory, the one that we associated with the $B$-model of complex
structure $I$. Making or not making a $T$-duality on $A_1$ or
$A_3$ does not affect the fact that $\phi$ obeys the $B$-model
condition $[Q,\phi]=0$. But in the absence of any $T$-duality,
the conditions obeyed by $A_1$ and $A_3$ are $A$-model conditions,
not $B$-model conditions.
Indeed,
in our derivation, we used the fact that $\partial_0A_1+\partial_2A_3$ is $Q$-exact in the
effective two-dimensional theory;
similarly, the same is true of $\partial_0 A_3-\partial_2 A_1$. We can combine these
statements into the assertion that $(\partial_0+i\partial_2)(A_3+iA_1)$ is $Q$-exact.
In the complex structure on ${\mathbb R}\times I$ in which $z=x^0+ix^2$ is holomorphic, this says
that $A_3+iA_1$ is holomorphic modulo $\{Q,\cdot\}$.
Thus the model under discussion, in terms of the obvious variables
without any $T$-duality, is from the point of view of the
supercharge $Q$ a mixed $AB$ model, with $B$-model conditions on
$\phi$ and $A$-model conditions on $A_3+iA_1$. A similar story
holds if $Q$ is replaced by one of the other supercharges
considered above.
\subsubsection{More On Compactification}\label{morec}
\def{\eusm C}{{\eusm C}}
We will conclude this discussion with a few more observations
about compactification of ${\mathcal N}=2$ gauge theories from four to two
dimensions. The goal is to review some points made in
\cite{HM,BJV} and derive some formulas that will be used later.
(See the appendix for a much more complete treatment.)
We will here consider only compactification of a $U(1)$ vector
multiplet on a rectangular torus $T^2$. We write ${\eusm C},{\eusm C}'$ for
the circumferences of the two circles. Also, for simplicity, we
take the four-dimensional $\theta$-angle to vanish.
The scalar $\phi$ in the vector multiplet reduces to a scalar in two dimensions that we denote
as $a$. Its kinetic energy is
\begin{equation}\label{boken} \frac{ {\eusm C} {\eusm C}'}{e^2}\int {\mathrm d}^2x\,|\nabla a|^2.\end{equation}
The components of the gauge field along $T^2$ have zero modes that reduce in two dimensions
to angle-valued field $b,b'$. Their kinetic energy is
\begin{equation}\label{zoken}\frac{1}{e^2}\int {\mathrm d}^2x\,\left(\frac{ {\eusm C}'}{ {\eusm C}}|\nabla b|^2+
\frac{ {\eusm C}}{ {\eusm C}'}|\nabla b'|^2\right).\end{equation} However, we
want a description obtained by $T$-duality on one of these
scalars. Making a $T$-duality that replaces $b'$ by another
angle-valued field $\varrho$, (\ref{zoken}) is replaced by
\begin{equation}\label{token}\frac{ {\eusm C}'}{ {\eusm C}}\int {\mathrm d}^2x\,\left(\frac{1}{e^2}|\nabla b|^2+
\frac{e^2}{16\pi^2}|\nabla \varrho|^2\right).\end{equation}
In this simple model, the angles $b,\varrho$ parametrize the
fibers of the Hitchin fibration. Clearly, the area of a fiber
$\cmmib F$ of the Hitchin fibration is independent of $e^2$ and is
proportional to $ {\eusm C}'/ {\eusm C}$. So $ {\eusm C}'/ {\eusm C}$ is a Kahler
parameter, as in \cite{HM,BJV}. If we rescale $a$ to $\widetilde a=a
{\eusm C}/e$, then the metric on ${\mathcal M}_H$ become $ {\eusm C}'/ {\eusm C}$ times a
metric that depends only on $e$ and not on $ {\eusm C}$ or $ {\eusm C}'$:
\begin{equation}\label{zongo}{\mathrm d} s^2=\frac{ {\eusm C}'}{ {\eusm C}}\left(|{\mathrm d} \widetilde a|^2+\frac{1}{e^2}{\mathrm d} b^2
+\frac{e^2}{16\pi^2}{\mathrm d} \varrho^2\right).\end{equation}
Conversely, the complex structure of $\cmmib F$ is determined by
$e^2$ and is independent of $ {\eusm C}$ and $ {\eusm C}'$. In fact, the
$\tau$ parameter of $\cmmib F$ is that of the underlying gauge
theory. If one thinks of the $U(1)$ vector multiplet as arising
from compactification of the abelian version of the
six-dimensional $(0,2)$ theory on a two-torus $C$, then ${\mathcal M}_H$ is
the moduli space of $U(1)$ Higgs bundles on $C$. This picture
extends to more interesting examples with $U(1)$ replaced by a
nonabelian group and $C$ by a more general Riemann surface. The
effective two-dimensional description in an example of this type
is obtained by compactifying the six-dimensional $(0,2)$ theory to
two dimensions on $T^2\times C$. If one compactifies first on
$C$, one gets a generalized quiver theory \cite{DG} in four
dimensions in which the complex structure of $C$ is encoded in the
gauge coupling parameters (generalizing what we called $e^2$ in
the abelian theory). If one compactifies first on $T^2$, one gets
${\mathcal N}=4$ super Yang-Mills theory with coupling parameter
$\tau=i{\eusm C}_1/{\eusm C}_2$. Further compactification on $C$ gives the
situation studied in \cite{KW}: supersymmetric vacua correspond to
Higgs bundles on $C$, so the moduli space of vacua is ${\mathcal M}_H$, and
the metric on ${\mathcal M}_H$ is scaled by ${\mathrm
{Im}}\,\tau={\eusm C}_1/{\eusm C}_2$, generalizing the abelian result of
(\ref{zongo}).
Since the metric has a factor of ${\eusm C}'/{\eusm C}$, the Kahler forms all
contain this factor as well. Writing
$(\omega_I^*,\omega_J^*,\omega_K^*)$ for the Kahler forms computed
relative to the metric (\ref{zongo}), these Kahler forms are
${\eusm C}'/{\eusm C}$ times the standard ones defined in eqns. (\ref{zobox}),
(\ref{obox}):
\begin{equation}\label{retso}(\omega_I^*,\omega_J^*,\omega_K^*)
=\frac{{\eusm C}'}{{\eusm C}}(\omega_I,\omega_J,\omega_K).\end{equation}
Of course, this factor will also appear in the holomorphic two-forms. For example, the holomorphic
two-form in complex structure $I$ is
\begin{equation}\label{metso}\Omega_I^*=
\omega_J^*+i\omega_K^*=\frac{{\eusm C}'}{{\eusm C}}\left(\omega_J+i\omega_K\right).\end{equation}
In fact, $\Omega_I^*=(e\,{\eusm C}'/2\pi{\eusm C} ){\mathrm d} \widetilde
a\wedge({\mathrm d}\varrho-(4\pi i/e^2){\mathrm d} b)=({\eusm C}'/2\pi){\mathrm d} a\wedge({\mathrm d}
\varrho-(4\pi i/e^2){\mathrm d} b)$. It is convenient to introduce
$a_D=\tau a$, with\footnote{Here $\tau$ is imaginary since we took
the four-dimensional $\theta$-angle to vanish. For a derivation
including the $\theta$-angle, see eqns. (2.34) and (3.16) of
\cite{GMN}.} $\tau=4\pi i/e^2$, in terms of which
\begin{equation}\label{poln}\Omega_I^*=\frac{{\eusm C}'}{2\pi}
\bigl({\mathrm d} a\wedge{\mathrm d}\varrho -{\mathrm d} a_D\wedge {\mathrm d} b\bigr).\end{equation}
The expression in parentheses is a standard formula for the
holomorphic two-form of the total space of the Seiberg-Witten
fibration over the Coulomb branch. (This holomorphic two-form is
unaffected by compactification.) The main purpose of computing
$\Omega_I^*$ here has been to explain the factor of ${\eusm C}'$.
The formula (\ref{poln}) has an analog for the general case of a
semi-simple gauge group of rank $r$. Pick a local description of
the Coulomb branch in terms of $r$ vector multiplets with scalar
components $a_i$ and dual scalar fields $a_D^i=\partial{\mathcal
F}/\partial a_i$ (here $\mathcal F$ is the prepotential of the
gauge theory). The gauge field of the $i^{th}$ vector field gives
two angle-valued fields $\varrho^i$ and $b_i$ (as in the $U(1)$
case, $b_i$ is a holonomy and $\varrho_i$ is the dual of a
holonomy). The analog of (\ref{poln}) is then
\begin{equation}\label{zpoln}\Omega_I^*=\frac{{\eusm C}'}{2\pi}\sum_i
\bigl({\mathrm d} a_i\wedge{\mathrm d}\varrho^- -{\mathrm d} a_D^i\wedge {\mathrm d}
b_i\bigr).\end{equation} Because it requires the choice of a
duality frame, this formula is valid only locally on the Coulomb
branch. Given the choice of a duality frame, the formula can be
derived by the same steps as in the $U(1)$ case, starting with the
low energy effective action on the Coulomb branch.
\subsection{Rethinking The $\Omega$-Deformation}\label{rethink}
By now, we have learned that the brane associated with the tip of the cigar manifold $D_R$
is a brane of type $(B,B,B)$ whose support is all of ${\mathcal M}_H$. On the other hand, we know
from section \ref{coiso} that this is not what we need in order to generate the quantization of
a real slice in ${\mathcal M}_H$. What we need is a brane of support ${\mathcal M}_H$ and of type $(A,B,A)$.
It is also clear from \cite{NS2} what we need to do in order to get the desired result: we need
to implement an $\Omega$-deformation -- that is, we need to replace $D_R$ by its
$\Omega$-deformed version
$D_{R,\varepsilon}$, where $\varepsilon$ is the parameter of
the $\Omega$-deformation.
\subsubsection{The Standard Formulation}\label{standard}
First, let us recall a standard way of introducing the
$\Omega$-deformation. One uses the fact that the scalar fields
$A_4$ and $A_5$ can be viewed as components of the gauge field
$A_I$ in two extra dimensions. For simplicity, let us focus on
$A_4$. We think of $x^4$ as an angular variable. Instead of taking
spacetime to be a product ${\mathbb R}^4\times S^1$, with the second factor
parametrized by $x^4$, we take it to be an ${\mathbb R}^4$ bundle over
$S^1$, with the monodromy around $S^1$ being an element $g\in
SO(4)$. The monodromy action on ${\mathbb R}^4$ can be accompanied by an
$R$-symmetry transformation $r\in SU(2)_R$ acting on the fermions,
and then the unbroken supersymmetries are those that are invariant
under the product $gr$. For our present purposes, we take $g$ to
rotate a two-plane in ${\mathbb R}^4$, so $g$ actually lies in an $SO(2)$
subgroup of $SO(4)$. The element $r$ is chosen in the usual way
so that the product $gr$ preserves one-half of the supersymmetry.
Now instead of taking the metric on ${\mathbb R}^4\times S^1$ to be a
simple product, we consider a fiber bundle metric in which the
monodromy around $S^1$ is the element $gr\in SO(4)\times SU(2)_R$
(in other words, the geometric monodromy is $g$, and we also make
an $R$-symmetry twist by $r$). In formulas, the metric is
\begin{equation}\label{frog}{\mathrm d} s^2=\sum_{\mu=0}^3({\mathrm d} x^\mu-\varepsilon V^\mu{\mathrm d} x^4)^2 +({\mathrm d} x^4)^2,\end{equation}
The $\Omega$-deformed four-dimensional theory is defined by starting with five-dimensional
super Yang-Mills theory in this spacetime, and then taking the fields to be independent of $x^4$.
\def{\mathcal W}{{\mathcal W}}
The $\Omega$-deformation has its name because it actually is a
deformation of the action. A basis of orthornormal vector fields
for the metric (\ref{frog}) is given by $u_\mu=\partial/\partial
x^\mu$, $\mu=0,\dots,3$, $u_4=\partial/\partial x^4 +\varepsilon
V^\mu\partial/\partial x^\mu$. The only one that is unusual is
$u_4$, and the extra term in $u_4$ means that the
$\varepsilon$-dependence of the Lagrangian can be computed roughly
by a substitution $A_4\to A_4+\varepsilon V^\mu D_\mu$. More
precisely, the deformation of the bosonic part of the action can
be computed by the following substitutions
\begin{align}\label{helpo}[A_4,A_5] & \to [A_4,A_5]+\varepsilon V^\mu D_\mu A_5\cr
D_\nu A_4& \to D_\nu A_4+\varepsilon V^\mu F_{\nu\mu}.\end{align}
The Yukawa couplings containing $A_4$ are modified in a similar way.
The $\Omega$-deformation preserves whatever supersymmetry commutes
with $gr$. However, the supersymmetry algebra is modified. The
reason for this is that usually the supersymmetry algebra of
${\mathcal N}=2$ super Yang-Mills theory closes modulo gauge transformations
generated by $A_4$ and $A_5$. But in the present situation, $A_4$
is effectively replaced by $A_4+\varepsilon V^\mu D_\mu$, and so
wherever a gauge transformation generated by $A_4$ would usually
appear, there is now an additional term that is $\varepsilon$
times the conserved charge ${\mathcal W}_V$ associated to the Killing vector
field $ V^\mu$. The most important special case \cite{Nek}
concerns the supercharge $Q$ that is associated to Donaldson
theory and the counting of instantons. In the undeformed theory,
its square is a gauge transformation; after $\Omega$-deformation,
it obeys $Q^2=\varepsilon {\mathcal W}_V$ modulo a gauge transformation.
\subsubsection{An Alternative Description}\label{alternative}
In constructing the $\Omega$-deformation,
we can replace ${\mathbb R}^4\times S^1$ by $M\times S^1$, where $M$ is any Riemannian
four-manifold and $g$ is an isometry of $M$. Our application will be to the case that
$M={\mathbb R}\times S^1\times D_R$, with $g$ a rotation of the cigar metric $D_R$ that
leaves fixed the tip of the cigar.
In its asymptotic region, $D_R$ is simply a product $I\times \widetilde S^1$, with the circumference
of $\widetilde S^1$ being $2\pi \rho$. Rotations of $D_R$
act in the asymptotic region by rotations of $\widetilde S^1$. Such rotations of the flat metric
${\mathbb R}\times S^1\times I\times \widetilde S^1$ preserve all supersymmetry, and are not accompanied
by an $R$-symmetry transformation.\footnote{This may be understood as follows.
To preserve half of the supersymmetry on ${\mathbb R}\times S^1\times D_R$, in the absence of
the $\Omega$-deformation, one twists the fermions, changing their spins. This
modifies how they transform under rigid rotations of $D_R$ in the curved region near
the tip of the cigar, but not in the asymptotic region.}
In principle, we could now proceed to study the claim of \cite{NS2} in the context of the
$\Omega$-deformed theory. The only problem is that the $\Omega$-deformed theory
is a different theory from ${\mathcal N}=2$ super Yang-Mills theory, and we have much less experience
with its dynamics.
Here we will avoid having to understand the dynamics of $\Omega$-deformed
theories, via the following device. We will show that if $M$ is of the form
$W\times\widetilde S^1$, with a product metric, then the $\Omega$-deformation associated
to a rotation of $\widetilde S^1$ can be eliminated by a change
of variables. Of course, our $M$ is of the stated form.
We proceed as follows. In the product situation, the $\Omega$-deformation can be described
formally as a substitution
\begin{equation}\label{turkey} A_4\to \widetilde A_4=A_4+\varepsilon\rho\frac{D}{Dx^3}.
\end{equation}
(The reason for the factor of $\rho$ is that, as $x^3$ has period
$2\pi\rho$, the normalized generator of the rotation of $\widetilde
S^1$ is the vector field $\rho\partial/\partial x^3$.) This is a
substitution only in a very formal sense; because the right hand
side of (\ref{turkey}) is a differential operator of degree 1
rather than a field, this operation does not make sense as a
change of variables in any standard sense. However, the
substitution (\ref{turkey}) does make sense as a formal device to
generate a deformation of the action of ${\mathcal N}=2$ super Yang-Mills
theory. The reason for this is that $A_4$ enters the action only
via commutators such as $[A_4,A_5]$ and covariant derivatives
$D_\nu A_4$. Under the substitution (\ref{turkey}), these
transform in a sensible way as indicated in eqn. (\ref{helpo}).
Now, consider the transformation
\begin{equation}\label{durkey}\frac{D}{Dx ^3}\to \frac{D}{Dx ^3}-\varepsilon \rho A_4.\end{equation}
Though superficially similar, this does make sense as an ordinary change of variables
in the quantum field theory, since it is equivalent to
\begin{equation}\label{yurkey}A_3\to \widetilde A_3=A_3-\varepsilon\rho A_4, \end{equation}
and the right hand is a field rather than a differential operator of positive degree.
Since the distinction between the formal operation (\ref{turkey})
and an ordinary change of variables such as (\ref{durkey}) will be
important, we will belabor the point slightly. If the ordinary
change of variables (\ref{durkey}) is applied to a given quantum
field theory, we simply get an equivalent theory written in terms
of a different set of variables. By contrast, the operation
(\ref{turkey}) is not a change of variables in any ordinary sense,
and when it is applied to a theory to which it can be applied, it
gives a theory that in general is inequivalent. In general, the
$\Omega$-deformation really is a non-trivial deformation. However,
we will show here that in the special case of a product $W\times
\widetilde S^1$ and a deformation involving a rotation of the second
factor, the $\Omega$-deformation is trivial, in the sense that it
can be removed by an ordinary change of variables. The
identification of the deformed theory with the undeformed one
will, however, involve a non-trivial transformation of the
observables, and that is why the analysis will lead to something
useful.
The combination of (\ref{turkey}) and (\ref{durkey}) can be
viewed, formally, to first order in $\varepsilon$, as a rotation
of the $34$ plane. As such, it leaves fixed, to first order, the
bosonic part of the Lagrangian that contains $A_3$ and $A_4$. For
example, the contributions
\begin{equation}\label{luckey} \int_{W\times \widetilde S^1}{\mathrm d}^4x\sum_{\mu=0,1,2}{\mathrm {Tr}}\,\,\left(F_{\mu 3}^2
+D_\mu A_4^2\right)\end{equation} are invariant to first
nontrivial order in $\varepsilon$ under the combination of the
formal operation (\ref{turkey}), which we implement via
(\ref{helpo}), and the actual change of variables (\ref{durkey}).
The same is true for the other bosonic terms involving $A_3$ and
$A_4$, namely
\begin{equation}\label{plucky}\int_{W\times\widetilde S^1} {\mathrm d}^4x \,{\mathrm {Tr}}\,\left((D_3A_5)^2+[A_4,A_5]^2\right)
\end{equation}
and
\begin{equation}\label{zucky}\int_{W\times\widetilde S^1}{\mathrm d}^4x \,{\mathrm {Tr}}\, \,(D_3A_4)^2.\end{equation}
To compensate for the $\Omega$-deformation of the terms in the
action involving fermions, we must make a similar rotation of the
fermions
\begin{equation}\label{fermrot}\Psi\to \exp\left(\frac{\varepsilon\rho}{2}\Gamma_{34}\right)\Psi.\end{equation}
The formulas we have considered so far compensate for the
$\Omega$-deformation only to first non-trivial order in
$\varepsilon$, because the formulas (\ref{turkey}) and
(\ref{durkey}) represent a rotation of the 34 plane only to that
order. (By contrast, for the fermions we have used an exact
rotation matrix in (\ref{fermrot}).) How can we improve our
formulas to represent a rotation of the 34 plane beyond first
order?
For $A_4$, it is fairly clear what to do. We compose the
$\Omega$-deformation with an ordinary change of variables $ A_4\to
\widehat A_4=A_4/\sqrt{1+\varepsilon^2\rho^2}$. So the modified $A_4$
becomes
\begin{equation}\label{lanky}\widehat A_4=\frac{1}{\sqrt{1+\varepsilon^2\rho^2}}
\left(A_4+\varepsilon\rho\, \frac{D}{Dx^3}\right).\end{equation}
Now it is more or less clear what the remaining formula for the
redefinition of $A_3$ ought to be. To get a rotation of the 34
plane, we would like the formula to be in some sense
\begin{equation}\label{zanky}\frac{D}{Dx^3}\to \frac{1}{\sqrt{1+\varepsilon^2\rho^2}}
\left( \frac{D}{Dx_3}-\varepsilon\rho A_4\right).\end{equation}
The only problem is to explain what this formula means, as we are not free to
rescale the derivative $\partial/\partial x^3$.
We should interpret rescaling of $D/Dx^3$ as rescaling of this
covariant derivative referred to an orthonormal frame, which will
result from rescaling of the metric in the $x^3$ direction.
Instead of saying that $x^3$ has period $2\pi\rho$, we write
$x^3=\rho w$ where $w$ is an ordinary angular variable of period
$2\pi$. Thus the metric on ${\mathbb R}\times S^1\times I\times \widetilde
S^1$ becomes
\begin{equation}\label{oggo} {\mathrm d} s^2 =\sum_{\mu=0}^2({\mathrm d}
x^\mu)^2+\rho^2\, {\mathrm d} w^2.\end{equation} Let us now rewrite the contribution (\ref{luckey}) to the action
making explicit the dependence on
$g_{ww}=\rho^2$ and also on the gauge coupling constant $e^2$:
\begin{equation}\label{tuckey}\frac{1}{e^2}\int_{W\times\widetilde
S^1} {\mathrm d}^3x\,{\mathrm d} w \sqrt{g_{ww}}
\sum_{\mu=0,1,2}{\mathrm {Tr}}\,\,\left(g_{ww}^{-1}F_{\mu w}^2 +D_\mu
A_4^2\right).\end{equation} For this to be invariant, we should
rotate $A_4$ into $g_{ww}^{-1/2}D/D w$, which is the covariant
derivative referred to an orthonormal frame. This will entail
rescaling of $g_{ww}=\rho^2$. At the same time we will rescale
$e^2$ to ensure that
\begin{equation}\label{gensu}\frac{\sqrt{g_{ww}}}{e^2}\end{equation}
remains fixed. Given this, a rotation of the $A_4-X$ plane, where
$X=g_{ww}^{-1/2}D/Dw$ (which in the undeformed theory is the
same as $ D/Dx^3$), will leave (\ref{tuckey}) invariant. The
equations for the rotation should thus be
\begin{align}\label{talign} A_4&=\frac{1}{\sqrt{1+\varepsilon^2\rho^2}}
\biggl(\widehat A_4-\varepsilon\rho\, \widehat X\biggr)\cr
X &=\frac{1}{\sqrt{1+\varepsilon^2\rho^2}} \biggl(\widehat
X+\varepsilon\rho\, \widehat A_4\biggr),\end{align} where it is
convenient to express the objects $A_4, X$ of the undeformed
description in terms of corresponding deformed objects $\widehat A_4$,
$\widehat X$. Interpreting $\widehat X$ as $\widehat g_{ww}^{-1/2}D/Dw$, we
require
\begin{equation}\label{gwelf}
\frac{1}{\sqrt{g_{ww}}}=\frac{1}{\sqrt{1+\varepsilon^2\rho^2}}\frac{1}{\sqrt{\widehat g_{ww}}}
\end{equation}
or more simply
\begin{equation}\label{onelf}
\widehat g_{ww}=\frac{g_{ww}}{1+\varepsilon^2\rho^2}=\frac{\rho^2}{1+\varepsilon^2\rho^2}.\end{equation}
To keep (\ref{gensu}) fixed, we also need to change the gauge
coupling so that $\sqrt{\widehat g_{ww}}/\widehat
e^2=\sqrt{g_{ww}}/{e^2}$, or
\begin{equation}\label{nelf}\widehat e ^2=
\frac{e^2}{\sqrt{1+\varepsilon^2\rho^2}}.\end{equation}
The parameter $\varepsilon$ is a physical parameter of the
$\Omega$-deformation, and for example it enters the results of
\cite{NS2} that we aim to understand. However, the asymptotic
radius $\rho$ in the $\Omega$-deformed description is not an
important parameter for the topological field theory observables
of interest.
Accordingly, we could proceed with any choice of $\rho$. However,
it turns out that we get a particularly simple description if we
take $\rho$ to be very large, $\rho>>1/\varepsilon$, so we will
consider this case first. (In fact, in discussions of the
$\Omega$-deformation, a flat metric on ${\mathbb R}^2$ is typically
assumed, rather than our cigar metric $D$. This corresponds to
taking $\rho=\infty$.)
The above formulas simplify in the limit $\rho\to\infty$. The
formulas (\ref{talign}) expressing the objects $X,A_4$ of the
$\Omega$-deformed description in terms of analogous objects $\widehat
X,\widehat A_4$ in an undeformed description reduce to
\begin{align}\label{zlign} A_4&=-\widehat X\cr
X& =\widehat A_4.\end{align}
Thus, the transformation from the $\Omega$-deformed description to
an equivalent undeformed description is a simple $\pi/2$ rotation
of the 34 plane. The limiting value of $\widehat g_{ww}$ is
$1/\varepsilon^2$, which means that in undeformed language, we
take the radius of $\widetilde S^1$ to be $1/\varepsilon$:
\begin{equation}\label{blign}
\widehat\rho=\frac{1}{\varepsilon}.\end{equation}
Finally, for $\rho\to\infty$, we have $\widehat e\to 0$ and the
theory becomes weakly coupled. Since we can make the coupling
arbitrarily weak by taking $\rho$ large, one may wonder how the
theory can do anything interesting at all. The answer to this
question is that our formulas such as (\ref{nelf}) relating the
$\Omega$-deformed theory to an ordinary one are only valid away
from the tip of the cigar. For $\rho\to\infty$, quantum effects
remain, but they are localized near the tip of the cigar. Though
our explanation here is different in detail, our conclusion is the
same as in \cite{Nek}: when the $\Omega$-deformation is defined
using a vector field $V$, quantum effects are localized near
zeroes of $V$, because the effective gauge coupling becomes small
far away from the zeroes.
Thus, as long as we consider quantities that do not depend on
$\rho$, so that we can take $\rho$ large, the $\Omega$-deformed
theory on ${\mathbb R}\times S^1\times D_R$ is equivalent, away from the
tip of $D_R$, to a weakly coupled and undeformed theory in which
the asymptotic radius of $\widetilde S^1$ is $1/\varepsilon$. However,
in going from the $\Omega$-deformed description to this weakly
coupled and undeformed description, physical observables undergo a
$\pi/2$ rotation in the tangent space to the 34 plane. Such a
rotation acts on fermions as multiplication by
$(1+\Gamma_{34})/\sqrt 2$, so that is the transformation of the
supersymmetry generator $\eta$ in comparing the deformed
description to the weakly coupled and undeformed one:
\begin{equation}\label{kelmit}\eta\to\frac{1+\Gamma_{34}}{\sqrt 2}\eta.\end{equation}
Conjugation by this matrix maps $\Gamma_3$ to $\Gamma_4$ and
$\Gamma_4$ to $-\Gamma_3$.
\subsection{The Deformed Brane}\label{deformed}
We aim to generalize our result of section \ref{undeformed} and
describe in two-dimensional terms the half-BPS brane
${\mathcal B}_\varepsilon$ that arises by compactification on
$D_{R,\varepsilon}$, with the $\Omega$-deformation. In
$\Omega$-deformed language, the supersymmetries preserved by this
brane are those whose generator obeys eqn. (\ref{zomely}):
\begin{equation}\label{umok}\left(\Gamma_{23}+\sigma_{23}\right)\eta=0.\end{equation}
Since the symmetry that we used in making the $\Omega$-deformation
commutes with the matrix on the left hand side of (\ref{umok}),
the same condition also characterizes the unbroken supersymmetries
of the $\Omega$-deformed brane. However, to describe the
$\Omega$-deformed brane in conventional language (in the simplest
case, the limit of large $\rho$), we must make pick a rotation of
the 34 plane. The rotation angle $\vartheta$ can be read off from
eqn. (\ref{talign}):
\begin{equation}\label{zerro}\cos\vartheta
=\frac{1}{\sqrt{1+\varepsilon^2\rho^2}}.\end{equation}
\subsubsection{The Case
$\varepsilon\rho\to\infty$}\label{largerho}
Neither the $\Omega$-deformation nor the change of variables that
maps us back to the undeformed theory affects the support of the
brane ${\mathcal B}_\varepsilon$ that comes from the tip of the cigar. So
${\mathcal B}_\varepsilon$ is a half-BPS brane whose support is all of
${\mathcal M}_H$, just like the brane ${\mathcal B}_*$ that arises in the absence of
the $\Omega$-deformation. However, the supersymmetry preserved by
${\mathcal B}_\varepsilon$ is different from that preserved by ${\mathcal B}_*$.
We first consider the simplest case, namely the limit of large
$\rho$. In this limit, $\vartheta=\pi/2$ and the necessary
rotation of supersymmetry generators in going from
$\Omega$-deformed variables to standard ones was given in eqn.
(\ref{kelmit}). The rotation in question just transforms
$\Gamma_3$ to $\Gamma_4$, so in a formalism in which the
Lagrangian away from the tip of the cigar is standard, the
supersymmetries preserved by the $\Omega$-deformed brane
${\mathcal B}_\varepsilon$ are characterized not by (\ref{umok}) but by
\begin{equation}\label{zumok}\left(\Gamma_{24}+\sigma_{23}\right)\eta=0.\end{equation}
Just as in section \ref{undeformed}, it is convenient to view
$M={\mathbb R}\times S^1\times D_{R,\varepsilon}$ as an asymptotic $T^2$
bundle over ${\mathbb R}\times I$. We want to understand what topological
field theory structures ${\mathcal B}_\varepsilon$ preserves in the
effective two-dimensional theory.
As in section \ref{undeformed}, any two-dimensional topological
field theory structure is associated to a supersymmetry generator
$\eta$ that (modulo a possible $SU(2)_R$ transformation) obeys
\begin{equation}\label{urft}\left(\Gamma_{02}+\sigma_{31}\right)\eta=0.\end{equation}
The conditions (\ref{zumok}) and (\ref{urft}) select a
two-dimensional space of $\eta$'s. Any non-zero $\eta$ in this
two-dimensional space determines a topological field theory, so
just like ${\mathcal B}_*$, ${\mathcal B}_{\varepsilon}$ is compatible with a family
of topological field theories parametrized by $\mathbb{CP}^1$. To
select a particular member of this family, we need to place an
additional condition on $\eta$; any condition will do.
By analogy with section \ref{undeformed}, we consider three simple
conditions. One of these will be
\begin{equation}\label{plumkin}\left(\Gamma_{23}+\sigma_{12}\right)\eta=0.\end{equation}
What we get with this choice is easily determined. The three
conditions (\ref{zumok}), (\ref{urft}), and (\ref{plumkin}) are
equivalent modulo an $SU(2)_R$ symmetry to the three conditions
used in section \ref{undeformed} to characterize the $B$-model in
complex structure $J$, namely eqns. (\ref{zomely}),
(\ref{tomely}), and (\ref{domely}). (To compare our current set
of three conditions to the previous three, one uses the fact that
if $A\eta=B\eta=0$, then $[A,B]\eta=0$, and one also makes an
$SU(2)_R$ transformation $(\sigma_1,\sigma_2,\sigma_3)\to
(\sigma_3,\sigma_2,-\sigma_1)$.) So the brane ${\mathcal B}_\varepsilon$ is
a $B$-brane in complex structure $J$.
That is the only complex structure in which ${\mathcal B}_\varepsilon$ is a
$B$-brane. It turns out that the other simple properties of
${\mathcal B}_\varepsilon$ are conveniently understood by imposing on $\eta$
the additional condition
\begin{equation}\label{guro}\left(\Gamma_{21}+\sigma_{12}\right)\eta=0\end{equation}
or
\begin{equation}\label{puro}\left(\Gamma_{25}+\sigma_{12}\right)\eta=0\end{equation}
together with (\ref{zumok}) and (\ref{urft}):
\begin{align}\label{holf}\left(\Gamma_{24}+\sigma_{23}\right)\eta&=0\cr
\left(\Gamma_{02}+\sigma_{31}\right)\eta&=0.\end{align}
Either condition (\ref{guro}) or (\ref{puro}) leads to an
$A$-model for the effective topological field theory on ${\mathbb R}\times
I$. For example, let us consider (\ref{puro}). We will show that
it leads to an $A$-model for the symplectic structure $\omega_I$.
This means that the fields $A_4-iA_5$ and $A_1+i\varrho$ that are
holomorphic in complex structure $I$ on ${\mathcal M}_H$ will both obey
$A$-model conditions; in other words, they depend holomorphically
on $z=x^0+ix^2$, modulo exact terms. We will show this by
repeating the analysis of section (\ref{twod}), but now, instead
of the $T$-duality $A_3\leftrightarrow\varrho$ converting a mixed
$AB$-model to a $B$-model, it will convert a slightly different
mixed $AB$-model to an $A$-model.
If one places all $\sigma$-matrices on the right in equations
(\ref{puro}) and (\ref{holf}), and then multiplies the left and
right hand sides of the three equations, one learns that
\begin{equation}\label{prof}
\Gamma_{0245}\eta=-\eta,\end{equation} as a result of which the
six-dimensional chirality condition implies also that
\begin{equation}\label{trof} \Gamma_{13}\eta=i\eta.\end{equation}
Now, consider the group $SO(4)_{0245}$ that rotates the tangent
spaces to the bosonic fields $A_0,A_2,A_4,A_5$. Decomposing its
double cover as $SU(2)_{\ell '}\times SU(2)_{r'}$, (\ref{prof})
implies that $\eta$ is invariant under $SU(2)_{\ell'}$, and eqns.
(\ref{puro}) and (\ref{holf}) further imply that $\eta$ is
invariant under a diagonal subgroup of $SU(2)_{r'}\times SU(2)_R$.
In short, $\eta$ is characterized by precisely the conditions that
characterize the supersymmetry generator of Donaldson theory,
modulo a rotation of the tangent space that exchanges $A_3$ with
$A_4$ and $A_1$ with $A_5$. This means that we can borrow the
analysis of section (\ref{twod}), except that the roles of
$A_1+iA_3$ and $A_4-iA_5$ are now reversed. In section
(\ref{twod}), $A_4-iA_5$ obeyed a $B$-model condition, and
$A_1+iA_3$ obeyed an $A$-model condition that was converted to a
$B$-model condition by the $T$-duality
$A_3\leftrightarrow\varrho$. So now, $A_4-iA_5$ obeys an
$A$-model condition, and $A_1+iA_3$ obeys a $B$-model condition
that is converted to an $A$-model condition by the $T$-duality on
$A_3$.
We can analyze the consequence of eqn. (\ref{guro}) in the same
way. This equation differs from (\ref{puro}) by the exchange
$A_1\leftrightarrow A_5$. So now, mimicking the last sentence of
the last paragraph, $A_4-iA_1$ obeys an $A$-model condition, and
$A_5+iA_3$ obeys a $B$-model condition that is converted to an
$A$-model condition by the $T$-duality on $A_3$. So we get an
$A$-model in which $A_4-iA_1$ and $A_5+i\varrho$ are holomorphic
modulo exact terms.
Given our conventions for what is $J$ and what is $K$, the
functions $A_4-iA_1$ and $A_5+i\varrho$ are holomorphic in
complex structure $K$, and this means that the $A$-model that we
have just arrived at is the $A$-model of symplectic form
$\omega_K$.
In summary, the brane ${\mathcal B}_\varepsilon$ is -- in the limit of large
$\rho\varepsilon$ -- a brane of type $(A,B,A)$. For understanding
the results of \cite{NS2}, the most important property is that it
is an $A$-brane of type $\omega_K$. Indeed, as we see if we
rotate back to the deformed language, the topological supercharge
of the $A$-model of type $\omega_K$ is the one that corresponds to
the usual supercharge of the $\Omega$-deformed theory (or
equivalently of Donaldson theory). So this is the supercharge
that we must use to make contact with the results of \cite{NS2}.
It is also the right one for a different reason described in
section \ref{coiso}: the coisotropic brane with Chan-Paton
curvature $F=\omega_J$ in the $A$-model with symplectic structure
$\omega_K$ is the right tool if we want to study quantization of a
real section of the integrable system ${\mathcal M}_H$.
By contrast, in section \ref{confblocks}, where we make contact
with the relation \cite{AGT} of four-dimensional gauge theory to
two-dimensional Liouville theory, the important fact will be that
the same brane ${\mathcal B}_\varepsilon$ is an $A$-brane for the symplectic
structure $\omega_I$.
\subsubsection{The General Case}\label{gencase}
We have by now interpreted the supersymmetry preserved by the tip
of the cigar in the $\Omega$-deformed theory in undeformed
language for $\varepsilon\rho>>1$. However, since the
$\Omega$-deformation is most directly understood as a deformation
of the standard theory, with $\varepsilon$ as a small deformation
parameter, one would naturally like to analyze the opposite limit
$\varepsilon\rho<<1$. In fact, it is not hard to analyze the
general case using tools described in section \ref{little}.
We recall that a general topological twist in two-dimensions is
determined by a pair of complex structures $({\mathcal J}_+,{\mathcal J}_-)\in
{\Bbb{CP}}^1_+\times {\Bbb{CP}}^1_-$. A brane of type $(B,B,B)$ is
characterized by ${\mathcal J}_-={\mathcal J}_+$. On the other hand, as explained in
section \ref{genhalf}, a brane of type $(A,B,A)$ is characterized
by ${\mathcal J}_-=h{\mathcal J}_+$, where $h$ is a $\pi$ rotation around the $J$
axis.
We get from one to the other using the rotation (\ref{zlign}). We
view this rotation as an element $g$ of a group $SO(4)_{1345}$
that acts on the tangent space to dimensions 1345. As such, $g$
is a $\pi/2$ rotation of the $34$ plane. The double cover of
$SO(4)_{1345}$ is the group ${\Gamma}=SU(2)_+\times SU(2)_-$, introduced
in section \ref{rotgroup}, that acts on ${\Bbb{CP}}^1_+\times{\Bbb{CP}}^1_-$. As
a $\pi/2$ rotation of one plane, $g$ corresponds to a pair
$(g_+,g_-)\in SU(2)_+\times SU(2)_-$, where each of $g_\pm$
corresponds to a $\pi/2$ rotation in $SO(3)_\pm=SU(2)_\pm/{\Bbb Z}_2$.
The $SU(2)$ element $h$ of the last paragraph is then
$h=g_-^{-1}g_+$, as noted in section \ref{rotgroup}. Given that
$h$ is a $\pi$ rotation around the $J$ axis, evidently $g_+$ and
$g_-$ are rotations around the $J$ axis by respective angles
$\pi/2$ and $-\pi/2$.
Next let us consider the case of general $\varepsilon\rho$. $g$
is now a rotation of the 34 plane by a more general angle
$\vartheta$. So $g_+$ and $g_-$ are rotations around the $J$ axis
by angles $\vartheta$ and $-\vartheta$, respectively.
Now, starting with the fact that in the undeformed theory the tip
of the cigar preserves supersymmetry of type $(B,B,B$), we can
understand the supersymmetry preserved by the brane
${\mathcal B}_\varepsilon$ for general $\varepsilon\rho$. First of all,
complex structure $J$ is invariant under rotations around the $J$
axis. So ${\mathcal B}_\varepsilon$ is a $B$-brane of type $J$ for all
$\varepsilon\rho$. One can verify this explicitly by observing
that the conditions (\ref{plumkin}) and (\ref{holf}) are invariant
under a rotation of the 34 plane, modulo an $SU(2)_R$
transformation generated by $\sigma_{31}$.
More pressing for our application is to interpret in undeformed
variables the usual supercharge of the $\Omega$-deformed theory,
which in $\Omega$-deformed variables corresponds to ${\mathcal J}_+={\mathcal J}_-=I$.
After rotating ${\mathcal J}_+$ by an angle $\vartheta$ and ${\mathcal J}_-$ by an
angle $-\vartheta$ in the $IK$ plane, they map to
\begin{align}\label{porz} {\mathcal J}_+& =I \cos\vartheta+ K\sin\vartheta \cr
{\mathcal J}_-&= I\cos\vartheta - K\sin\vartheta. \cr\end{align}
According to (\ref{golf}), the topological field theory with this
pair $({\mathcal J}_+,{\mathcal J}_-)$ is equivalent to an $A$-model with
\begin{align}\label{ogolf}\omega^{-1}&=\frac{1}{2}\left(\omega_+^{-1}-\omega_-^{-1}\right)
\cr \omega^{-1}B&=\frac{1}{2}({\mathcal J}_++{\mathcal J}_-).\end{align} Here, writing
$g$ for the hyper-Kahler metric on ${\mathcal M}_H$, we have
$\omega_\pm=g{\mathcal J}_\pm$.
To make (\ref{ogolf}) more explicit, we use
$\omega_+=\omega_I^*\cos\vartheta+\omega_K^*\sin\vartheta$,
$\omega_-=\omega_I^*\cos\vartheta-\omega_K^*\sin\vartheta$. (The
meaning of the asterisks was explained in eqn. (\ref{retso}):
$\omega_I^*$ and $\omega_K^*$ are the Kahler forms for the actual
Kahler metric on ${\mathcal M}_H$, which depends on the radii of the
compactification.) Evaluating (\ref{golf}), we find that the
symplectic form of the equivalent $A$-model is
\begin{equation}\label{dorf}\omega=\frac{\omega_K^*}{\sin\vartheta}.\end{equation}
So the supersymmetry is that of the $A$-model with symplectic
structure a multiple of $\omega_K$. The multiple is inessential as
explained in the discussion of eqn. (\ref{oplo}). The result
agrees, as expected, with what we found for large
$\varepsilon\rho$ in section (\ref{deformed}).
Eqn. (\ref{golf}) also tells us how to determine the $B$-field:
\begin{equation}\label{torf}\omega^{-1}B=I\cos\vartheta,\end{equation}
which implies that $B$ is a multiple of $\omega_J$.
\subsubsection{A Canonical Coisotropic Brane}\label{canco}
Since $\omega_J$ is cohomologically trivial in a large class of
models, as explained in section \ref{distinguished}, one may
wonder what we actually learn by determining $B$. The answer is
that we do not learn much until we also consider the couplings
generated at the tip of the cigar. Those couplings generate a
Chan-Paton curvature $F$ for the brane in the effective
two-dimensional description, and only the combination $F+B$ is
invariant under $B$-field gauge transformations.
In this paper, we will not attempt to compute the couplings at the
tip of the cigar. But we can make a simple observation. Let us
consider the limit in which $\varepsilon\rho$ is very small. We
approach this limit by keeping $\rho$ fixed and taking
$\varepsilon$ to zero. In this limit, the $\varepsilon$-dependent
corrections at the tip of the cigar are negligible, and $F$
vanishes. Since also $\vartheta\to 0$ for $\varepsilon\rho\to 0$,
eqn. (\ref{torf}) becomes $\omega^{-1}(F+B)=I$.
In other words, for $\varepsilon\rho$ very small, the brane
derived from the tip of the cigar approaches the coisotropic brane
of type $(A,B,A)$ in which the complex structure
${\mathcal I}=\omega^{-1}(F+B)$ is equal to $I$. Since
$\omega=\omega_K^*/\sin\vartheta$, this brane has
$F+B=\omega_J^*/\sin\vartheta$. It is convenient to evaluate this
using (\ref{zpoln}), using the fact that $\omega_J^*={\mathrm
{Re}}\,\Omega_I^*$, and identifying ${\eusm C}'$ as $2\pi\widehat\rho$
(which for small $\varepsilon\rho$ is the same as $2\pi\rho$). For
small $\varepsilon\rho$, we can replace $\sin\vartheta$ by
$\varepsilon\rho$, whereupon $\rho$ cancels out and we get
\begin{equation}\label{zum}F+B={\mathrm{Re}}\,\frac{1
}{\varepsilon}\sum_i\biggl({\mathrm d} a_i\wedge {\mathrm d}\varrho^i-{\mathrm d} a_D^i\wedge
{\mathrm d} b_i\biggr)\end{equation} along with
\begin{equation}\label{um}\omega={\mathrm{Im}}\,\frac{1
}{\varepsilon}\sum_i\biggl({\mathrm d} a_i\wedge {\mathrm d}\varrho^i-{\mathrm d} a_D^i\wedge
{\mathrm d} b_i\biggr).\end{equation}
Recalling the formula (\ref{poln}) for
$\Omega_I=\omega_J+i\omega_K$, it follows that at least for small
$\varepsilon$, the brane coming from the tip of the cigar is a
space-filling brane in which the complex structure
${\mathcal I}=\omega^{-1}(F+B)$ is equal to $\omega_K^{-1}\omega_J=I$. Since
$\varepsilon\rho$ is small in this computation,
the difference
between the gauge couplings $e$ and $\widehat e$ in $\Omega$-deformed
and undeformed variables is not important and $I$ is simply the
complex structure in which the Hitchin fibration of the moduli
space ${\mathcal M}_H$ of vacua is holomorphic.
In the case of a generalized quiver theory \cite{DG} that arises
by compactification from six dimensions on a Riemann surface $C$,
the brane is the canonical coisotropic brane of type $(A,B,A)$
that is naturally defined using the complex structure on $C$ that
prevails at the tip of the cigar -- the one that is used in the
$\Omega$-deformed description. This is the right brane for any
$\rho$, since $\rho$ is irrelevant in the twisted topological
field theory.
Changing $\rho$ changes $\widehat e$ and therefore changes the
effective complex structure on $C$. However, the brane that we
have found makes sense for any choice of the complex structure on
$C$, because the concept of a brane of type $(A,B,A)$ is
independent of that complex structure. This was explained in
section \ref{distinguished}.
Our determination that the complex structure ${\mathcal I}=\omega^{-1}(F+B)$
is equal to $I$ can be tested in the following way. Let
$Q_\varepsilon$ be the topological supercharge of the
$\Omega$-deformed theory. For $\varepsilon=0$, away from the tip
of the cigar, there are no non-trivial $Q_\varepsilon$-invariant
local operators. But at the tip of the cigar, $Q_\varepsilon$
reduces to the topological supercharge $Q$ of Donaldson theory,
and its cohomology in the space of gauge-invariant local operators
is generated by the gauge-invariant polynomials ${\mathrm {Tr}}\,\,\phi^n$ in
the scalar field $\phi$; these are the usual local operators in
Donaldson theory. In the $\sigma$-model with target ${\mathcal M}_H$, the
gauge-invariant polynomials in $\phi$ become the holomorphic
functions on the base of the Hitchin fibration (and thus, the
commuting Hamiltonians of the integrable system). Let us compare
this to the answer in the $A$-model with symplectic structure
$\omega$. In bulk, the $A$-model admits no local observables of
ghost number zero; however, on a boundary labeled by a
space-filling coisotropic brane with complex structure ${\mathcal I}$, the
observables are the holomorphic functions in that complex
structure. Thus, the complex structure ${\mathcal I}$ on ${\mathcal M}_H$ must be one
in which the Hitchin fibration $\pi:{\mathcal M}_H\to\cmmib B$ is
holomorphic -- so that the holomorphic functions on $\cmmib B$
pull back to holomorphic functions on ${\mathcal M}_H$. Our result ${\mathcal I}=I$
is consistent with this, since $I$ has the required property.
For this reason and because the deformation theory of coisotropic
branes is rather rigid, we believe that the result ${\mathcal I}=I$ or
equivalently the determination (\ref{zum}) of $F+B $ is exact,
even though our derivation was only valid for small $\varepsilon$.
\subsection{Boundary Conditions At The Far End}\label{farend}
Hitherto, in the effective field theory on ${\mathbb R}\times I$, we have
analyzed a distinguished brane ${\mathcal B}_\varepsilon$ that arises from
the tip of the cigar. It is a space-filling coisotropic brane in
an $A$-model with target ${\mathcal M}_H$. The symplectic form of the
$A$-model is a multiple of $\omega_K$, and the gauge-invariant
field $F+B$ of the brane is the corresponding multiple of
$\omega_J$.
By contrast, what happens at the other end of the cigar is largely
up to us. We can pick any $A$-model boundary condition we want at
the second boundary $r=R$ of ${\mathbb R}\times S^1\times D_R$. If the
aim is to study quantization of a real section of ${\mathcal M}_H$, we
should pick a boundary condition that determines a Lagrangian
$A$-brane with support $L\subset {\mathcal M}_H$. Moreover, while
Lagrangian for $\omega_K$, $L$ should be symplectic for
$\omega_J$. We pick a flat line bundle ${\mathcal S}$ over $L$, and write
${\mathcal B}_L$ for the corresponding $A$-brane. The space of
$({\mathcal B}_\varepsilon,{\mathcal B}_L)$ strings will be a quantization of $L$ with
symplectic form $F+B$.
To orient ourselves, we will begin by constructing some choices of
$L$ directly by hand, before describing them in four-dimensional
gauge theory language. Also, we will begin with the free $U(1)$
vector multiplet, previously considered in section \ref{morec}.
We will consider two simple choices of Lagrangian submanifold.
Since the symplectic form in the abelian case is
\begin{equation}\label{umbo}\omega={\mathrm{Im}}\,\frac{1
}{\varepsilon}\biggl({\mathrm d} a\wedge {\mathrm d}\varrho-{\mathrm d} a_D\wedge
{\mathrm d} b\biggr),\end{equation} we can define a Lagrangian
submanifold $L_1$ by\footnote{Throughout our derivation, we have
taken $\varepsilon$ to be real. However, a generalization to
complex $\varepsilon$ may be trivially made by a $U(1)_R$ rotation
(which may also act on bare masses and scale parameters of the
gauge theory). In this generalization, eqn. (\ref{umbo}) is still
valid, and the second conditions in (\ref{krumbo}) and
(\ref{rumbo}) should read ${\mathrm{Im}}\,(a_D/\varepsilon)=0$ and
${\mathrm{Im}}\,(a/\varepsilon)=0$. }
\begin{equation}\label{krumbo}\varrho=0={\mathrm{Im}}\,a_D.\end{equation}
And we can define a second Lagrangian submanifold $L_2$ by
\begin{equation}\label{rumbo} b=0={\mathrm
{Im}} \,a.\end{equation}
Restricted to $L_1$, we have $F+B={\mathrm d}\lambda$ with
\begin{equation}\label{tumbo}\lambda=-\frac{{\mathrm{Re}}\,a_D}{\varepsilon}\wedge {\mathrm d} b. \end{equation}
In the WKB approximation, quantum states
correspond to values of $\mathrm{Re}\,a_D$ for which
$\oint\lambda$ is an integer multiple of $2\pi$. Since $ b$ is an
angular variable, this condition says that
$\mathrm{Re}\,a_D/\varepsilon$ should be an integer. We also set
$\mathrm{Im}\,a_D/\varepsilon$ to zero as part of the definition
of $L_1$. The two conditions combine to
\begin{equation}\label{porf}
\frac{a_D}{\varepsilon}\in{\Bbb Z},\end{equation} or equivalently
\begin{equation}\label{uporf}\exp(2\pi i
a_D/\varepsilon)=1.\end{equation} Both the result and also the
fact that the real and imaginary parts of $a_D$ have gone their
separate ways in the derivation are hopefully reminiscent of
section \ref{linko}. In the case of the free abelian vector
multiplet, the WKB approximation is valid.
We can quantize $L_2$ similarly. In this case, $F+B={\mathrm d}\lambda$
with
\begin{equation}\label{phumbo}\lambda=\frac{{\mathrm
{Re}}\,a}{\varepsilon}\wedge{\mathrm d}\varrho.\end{equation} Quantization gives
\begin{equation}\label{zporf}
\frac{a}{\varepsilon}\in{\Bbb Z},\end{equation} or equivalently
\begin{equation}\label{zuporf}\exp(2\pi i
a/\varepsilon)=1.\end{equation}
How can we generalize this to the case of a non-abelian gauge
group? Given a choice of duality frame, the symplectic form is
the obvious generalization of (\ref{umbo}),
\begin{equation}\label{gumbo}\omega={\mathrm{Im}}\,\frac{1
}{\varepsilon}\sum_i\biggl({\mathrm d} a_i\wedge {\mathrm d}\varrho^i-{\mathrm d}
a_D^i\wedge {\mathrm d} b_i\biggr).\end{equation} So it looks like we can
define an analog of $L_1$ by setting $\varrho^i=0={\mathrm
{Im}}(a_D^i/\varepsilon)$. Similarly we can imitate the
definition of $L_2$. The only problem with these definitions is
that they only make sense locally, since they require a choice of
duality frame.
To find a definition that makes sense globally in the nonabelian
case, we need a better point of view. Let us first return to the
$U(1)$ problem and express the definitions of $L_1$ and $L_2$ in a
different way. For gauge group $U(1)$, the moduli space of vacua
is the moduli space ${\mathcal M}_H$ of $U(1)$ Higgs bundles on a Riemann
surface $T$ that is simply a two-torus parametrized by the angles
$\varrho$ and $b$. In complex structure $J$, ${\mathcal M}_H$ parametrizes
flat ${\Bbb C}^*$ bundles over $T$ (${\Bbb C}^*$ is the complexification of
$U(1)$). Such a flat bundle is labeled by its holonomies.
Choosing an $A$-cycle and a $B$-cycle that correspond to the
$\varrho$ and $b$ directions on $T$, the two holonomies are
$U=\exp( i(\varrho+i\,{\mathrm{Im}}\,a_D/\varepsilon))$ and
$V=\exp( i(b+i\,{\mathrm{Im}}\,a/\varepsilon))$. Thus, our two
Lagrangian submanifolds $L_1$ and $L_2$ can be described by the
conditions $U=1$ and $V=1$, respectively.
This definition can be adapted to the nonabelian case, if one
considers generalized quiver theories \cite{DG} which arise by
compactifying from six dimensions on a Riemann surface $C$. In
fact, there are multiple ways to do this.
One approach, assuming that $C$ has genus $g$ (and for brevity no
punctures), is to pick a set of $g$ $A$-cycles and $g$ $B$-cycles
on $C$, generating the first homology of $C$. Then one can define
a complex Lagrangian submanifold $L\subset{\mathcal M}_H$ for complex
structure $J$ by requiring the holonomies of the flat bundle
around all $A$-cycles (or around all $B$-cycles) to be trivial.
This can lead to a construction in which formulas like
(\ref{uporf}) and (\ref{zuporf}) are valid at least
asymptotically.
For $G=SU(2)$, so that $G_{\Bbb C}=SL(2,{\Bbb C})$, another construction of a
complex Lagrangian submanifold that in some sense generalizes what
we did for $U(1)$ is as follows. In relation to generalized
quiver gauge theories \cite{DG}, it is convenient to build $C$
from $3g-3$ cylinders or tubes that are glued together via
trinions or ``pairs of pants.'' Topologically, such a
construction gives us $3g-3$ circles, one for each cylinder. A
complex Lagrangian submanifold $L\subset{\mathcal M}_H$ can be defined by
requiring that the holonomy of the flat $SL(2,{\Bbb C})$ bundle around
each circle is unipotent, in other words conjugate to a triangular
matrix
\begin{equation}\label{morz} \begin{pmatrix}1 & \star \cr 0 &
1\end{pmatrix}.\end{equation} A
generalization of this Lagrangian submanifold for $G$ of higher
rank is not immediately apparent.
To learn more, we now consider the construction of branes on
${\mathcal M}_H$ of type $(A,B,A)$ from the point of view of
four-dimensional gauge theory.
\subsubsection{Gauge Theory Realization}\label{gaugereal}
To construct branes of type $(A,B,A)$ in gauge theory, we need
half-BPS boundary conditions that preserve those supersymmetries
whose generator $\eta$ obeys
\begin{equation}\label{orf}M\eta =\eta\end{equation}
or equivalently \begin{equation}\label{oorf}\overline\eta M=\overline
\eta,\end{equation}
where
\begin{equation}\label{torfox}M=\Gamma_{24}\sigma_{23}.\end{equation}
This choice of $M$ corresponds to the case
$\varepsilon\rho\to\infty$, so in this analysis we are restricted
to that case.
We will consider only the most elementary examples of half-BPS
boundary conditions. The full classification and analysis of
half-BPS boundary conditions with ${\mathcal N}=2$ supersymmetry is likely
to be quite rich, just as in the case of ${\mathcal N}=4$ supersymmetry
\cite{GW2}. Here we will only scratch the surface.
In any local boundary conditions, half of the fermion fields will
vanish at the boundary. The obvious condition that will
accomplish this, while preserving all the symmetries of the matrix
$M$, is to place on the fermions $\Psi$ in the vector multiplet a
condition analogous to (\ref{orf}) but possibly with the opposite
sign:
\begin{equation}\label{yorf}M\Psi =z\Psi,~~z=\pm 1.\end{equation}
For either choice of the sign $z$, we will describe a boundary
condition on the bosons in the vector multiplet that preserves the
supersymmetries with $M\eta=\eta$.
The condition that a boundary condition preserves supersymmetry is
that it ensures the vanishing of the normal component of the
supercurrent at the boundary. The supercurrent is
$J_I=\Gamma^{JK}\mathrm{Tr}\,F_{JK}\Gamma_I\Psi$, and so, bearing
in mind that in our notation the normal coordinate is $x^2$, the
condition that the normal part of $\overline\eta J_I$ vanishes at the
boundary is that at the boundary we must have
\begin{equation}\label{porft}\overline\eta
\mathrm{Tr}\,\Gamma^{JK}F_{JK}\Gamma_2\Psi = 0. \end{equation} To
ensure this condition, we must pick a judicious boundary condition
on the bosons $A_I$.
We find the condition we need by replacing $\overline\eta$ on the left
hand side of (\ref{porft}) by $\overline\eta M$, moving $M$ to the
right, and then using $M\Psi=z\Psi$. If $z=1$, then, since
$M\Gamma_2=-\Gamma_2M$, the condition that we want is that the
boundary values should obey
\begin{equation}\label{tommy}
[M,\Gamma^{JK}F_{JK}]=0.\end{equation} If instead we take $z=-1$,
then we need instead
\begin{equation}\label{zommy} \{M,\Gamma^{JK}F_{JK}\}=0\end{equation}
at the boundary.
To obey (\ref{tommy}), we impose Dirichlet boundary conditions on
$A_4$ and Neumann boundary conditions on the other
components\footnote{From this and the other statements below, we
omit $A_2$, the normal component of $A$, as it can be set to zero
near the boundary by a gauge transformation.} of $A$. (The former
implies that $F_{I4}=0$, $I\not=2$, while the latter implies
$F_{I2}=0$, $I\not=4$; together these imply (\ref{tommy}).) And
conversely, to obey (\ref{zommy}), by an argument similar to the
one just indicated, we place Neumann boundary conditions on $A_4$
and Dirichlet boundary conditions on the other components of $A$.
There is an important detail to mention here. There is never a
problem with Dirichlet boundary conditions on gauge fields, but in
four dimensions, Neumann boundary conditions on gauge fields are
only possible if the gauge theory $\theta$-angles vanish. (Neumann
boundary conditions in our notation would say that the boundary
value of $F_{2I}$ vanishes, $I=0,1,3$ while at non-zero $\theta$,
one must add to this a multiple of $\theta\epsilon_{IJK}F_{JK}$.
Here $I,J,K$ take values 013.) So at $\theta\not=0$, we are
limited to $z=-1$ and (\ref{zommy}).
We can make these two boundary conditions more concrete as
follows:
(I) In the first case (which as just explained requires the
$\theta$-angles to vanish), $A_4$ vanishes on the boundary and
$A_1,A_3,$ and $A_5$ do not.
(II) In our second case, $A_1$, $A_3$, and $A_5$ vanish on the
boundary and $A_4$ does not.
We have described these boundary conditions for vector multiplets,
but they extend to hypermultiplets. In the case of a generalized
quiver theory obtained by compactification from six dimensions on
a Riemann surface $C$, the gauge group is semi-simple rather than
simple. We pick the same type of boundary condition -- type I or
type II -- for each factor.
Now let us describe what these boundary conditions look like after
toroidal compactification to two dimensions. First we consider a
free vector multiplet with gauge group $U(1)$. As usual, we
$T$-dualize the holonomy of $A_3$ to a scalar $\varrho$.
$\varrho$ obeys Dirichlet boundary conditions if $A_3$ obeys
Neumann boundary conditions, and vice-versa.
So in case I, the fields that vanish at the boundary are
$\varrho$ and $A_4$. We can think of $A_4$ as $\mathrm{Im}
\,a_D$ where $a=A_4-iA_5$ and $a_D=(4\pi i/e^2)a$. (We recall
that for Type I, the gauge theory $\theta$-angle vanishes and so
the gauge coupling parameter $\tau$ reduces to $4\pi i/e^2$.)
Thus, the Type I boundary condition leads to a brane supported on
the Lagrangian submanifold $L_1$.
In case II, the fields that vanish on the boundary are $A_1$ and
$A_5$. In the two-dimensional description, the fields that vanish
on the boundary are $b$, which is the holonomy of $A_1$, and
${\mathrm{Im}}\,a$, which is a multiple of the holonomy of $A_5$.
So the Type II boundary condition leads to a brane supported on
the Lagrangian submanifold $L_2$.
What about the nonabelian case? We cannot so easily interpret the
boundary conditions I and II in the low energy theory. But we can
do so asymptotically on the Coulomb branch. Generically on the
Coulomb branch, the gauge group is broken to an abelian subgroup;
the only massless particles are $r$ vector multiplets, with $r$
the rank of the gauge group. Near infinity on the Coulomb branch
(and far from the locus on which additional massless particles
appear), an abelian treatment along the above lines is a good
approximation.
\def{\mathcal I}{{\mathrm{I}}}
\def{\mathrm{II}}{{\mathrm{II}}}
So the brane ${\mathcal B}_{\mathcal I}$ coming from a Type I boundary condition can
be described near infinity by the familiar conditions
$\varrho^i={\mathrm {Im}}\,a_D^i=0$. Similarly the brane
${\mathcal B}_{\mathrm{II}}$ coming from a Type II boundary condition can be described
near infinity by $b_i={\mathrm{Im}}\,a_i=0$.
There are two reasons that we cannot simply imitate the derivation
of eqns. (\ref{porf}) and (\ref{zporf}) and determine the spectrum
by setting $a_i/\varepsilon$ or $a_D^i/\varepsilon$ to integers.
First, the branes ${\mathcal B}_{\mathcal I}$ and ${\mathcal B}_{\mathrm{II}}$ are described only near
infinity by the conditions mentioned in the last paragraph.
Second, in general the WKB approximation, which was used to arrive
at (\ref{porf}) and (\ref{zporf}), is not exact.
In section \ref{wkb}, we have described a situation in which the
WKB approximation is exact. This involved a two-dimensional
theory with $(2,2)$ supersymmetry, formulated on ${\mathbb R}\times S^1$.
We have been studying here, after toroidal compactification from
four dimensions, a theory with $(4,4)$ supersymmetry on ${\mathbb R}\times
I$, with half-BPS boundary conditions at the ends. As far as we
know, the WKB approximation is not exact in this situation.
\subsubsection{Alternative Compactification To Two
Dimensions}\label{altcom}
However, we can look at the same problem in another way. Our
starting point has been a four-dimensional gauge theory on
${\mathbb R}\times S^1\times D_{R,\varepsilon}$, understood in terms of
$S^1\times \widetilde S^1$ compactification to two dimensions. This
leads to a problem on ${\mathbb R}\times I$ that we have by now discussed
at length.
However, a more obvious way to reduce the same problem to
two dimensions is to simply view it as a compactification on
$D_{R,\varepsilon}$ down to ${\mathbb R}\times S^1$. In this fashion, we
arrive at a $(2,2)$ theory on ${\mathbb R}\times S^1$.
Let us consider the two cases of a Type I or Type II boundary
condition. In the first case, all vector multiplets obey Neumann
boundary condition. A four-dimensional vector multiplet with
${\mathcal N}=2$ supersymmetry reduces to a two-dimensional vector multiplet
with $(2,2)$ supersymmetry. Four-dimensional hypermultiplets
reduce to two-dimensional chiral multiplets. The net effect is
that a four-dimensional generalized quiver theory reduces to a
two-dimensional theory of $(2,2)$ supersymmetry based on the same
generalized quiver.
This is precisely the setting for section \ref{wkb}, and the WKB
approximation to the quantization will be exact when expressed in
terms of the effective twisted chiral superpotential $\widetilde W$,
which in general will depend on both the scalar fields in the
vector multiplet and the deformation parameter $\varepsilon$. This twisted chiral
superpotential has been analyzed in \cite{Nek} and \cite{NS2}. It receives contributions from
perturbative effects and instanton effects near the tip of the
cigar. General arguments concerning the $\Omega$-deformation show
that all contributions come from the region near the tip. As far
as we know, the analysis of $\widetilde W$ may as well be carried out
in $\Omega$-deformed variables; we know of no advantage to the
rotation to undeformed variables that has been exploited in the
present paper.
Since the effective twisted chiral superpotential is the same as
the one used in \cite{NS2}, the spectrum is also the same. All
that we have gained by our approach here is the understanding of
why this spectrum can be understood as coming from the
quantization of a real slice of Hitchin's integrable system.
The case of a Type II boundary condition is a little different.
All gauge fields obey Dirichlet boundary conditions. This
completely breaks the gauge symmetry, so there are no gauge fields
in the effective description on ${\mathbb R}\times S^1$. Both vector
multiplets and hypermultiplets reduce to chiral multiplets in that
effective description.
It seems that the clearest picture emerges if we rotate back from
the ordinary variables that we have used in describing the
boundary conditions to $\Omega$-deformed variables. This has the
effect of exchanging $A_4$, which is a scalar field (the real part
of $\phi$) with $A_3$, the component of the gauge field around
$\widetilde S^1$, the boundary of $D_R$. So in this description, the
boundary condition does not constrain the holonomy of $A_3$ around
$\partial D_R$.
However, for the purposes of constructing a supersymmetric ground
state, the holonomy around $\partial D_R$ must vanish, or
supersymmetry will be violated by curvature in the interior of
$D_R$. Treating $A_3$ as a constant in the low energy theory, the
holonomy is $\exp(2\pi \rho A_3)$, where $\rho$ is the radius of
$\widetilde S^1$ in $\Omega$-deformed variables, so the condition is
\begin{equation}\label{kornz}\exp(2\pi\rho A_3)=1.\end{equation}
We want to rotate this condition back to undeformed variables. In
doing so, we take $\rho\to\infty$, $\widehat\rho\to 1/\varepsilon$,
since this choice was built into our construction of boundary
conditions in section \ref{gaugereal}. In the undeformed
variables, $\rho A_3$ can be replaced by $A_4/\varepsilon$. So the
condition becomes $\exp(2\pi A_4/\varepsilon)=1$. The boundary
condition also set $A_5=0$. We can combine the two statements to
\begin{equation}\label{zelk}\exp(2\pi
\phi/\varepsilon)=1.\end{equation}
In the low energy description, we rotate $\phi$ to a maximal torus and denote it as $a$.
Thus, (\ref{zelk}) seems to show that the quantization of $a$ for a Type II boundary
condition takes its most naive form, though this is not so for Type I.
\subsubsection{Eigenvalues And Opers}\label{eigand}
Finally, we will describe what may ultimately -- after some future
developments -- be the most powerful way to get detailed results
about quantization of these integrable systems. Something close to
what we will describe momentarily has actually been carried out in
the mathematical literature \cite{F1,FFR,F2,F3} in the context of
an integrable system known as the Gaudin model. (This work has
been done in a language much closer to geometric Langlands than to
gauge theory, and at the moment we do not know precisely how to
reformulate it in terms of gauge theory or even two-dimensional
sigma-models.)
Our starting point in this section has been to choose a Lorentz-invariant boundary condition
at the far end of $D_R$ that preserves supersymmetry of type $(A,B,A)$.
In a description that arises by $T$-duality on scalars that represent holonomies around
$\widetilde S^1$, this boundary condition determines a Lagrangian submanifold $L\subset {\mathcal M}_H$
and a corresponding Lagrangian brane ${\mathcal B}_L$. Quantization of $L$ is carried out
by taking the space of $({\mathcal B}_\epsilon,{\mathcal B}_L)$ strings.
However, at the far
end of $D_R$, the two circles $S^1$ and $\widetilde S^1$ are on an equivalent footing.
Hence, had we made the $T$-duality on scalars associated to $\widetilde S^1$ rather than
$S^1$, the same boundary condition would have given a very similar Lagrangian
submanifold $\widetilde L$ again of type $(A,B,A)$. Actually $\widetilde L$ is a Lagrangian submanifold
not of ${\mathcal M}_H$ but of a dual moduli space $\widetilde{\mathcal M}_H$ obtained from ${\mathcal M}_H$ by $T$-duality on the
fibers of the Hitchin fibration.\footnote{In the context of generalized quiver gauge theories,
if ${\mathcal M}_H$ is a moduli space of Higgs bundles on a Riemann surface $C$ with gauge
group $G$, then \cite{HT} $\widetilde {\mathcal M}_H$ is a corresponding moduli space of Higgs bundles on
$C$ with gauge group $G^\vee$, the group dual to $G$. If $G$ is simply-laced, which is the
case that arises most simply by reduction from the $(0,2)$ model in six dimensions,
then $G$ and $G^\vee$ have the same universal cover, and so do ${\mathcal M}_H$ and $\widetilde{\mathcal M}_H$.
In this situation, $L$ and $\widetilde L$ should be equivalent if lifted to the universal cover.}
What happens to the brane ${\mathcal B}_\varepsilon$ if we use a description
obtained by $T$-duality on scalars coming from holonomy around
$S^1$ rather than $\widetilde S^1$? In this case, ${\mathcal B}_\varepsilon$ is
replaced by a new brane, also of type $(A,B,A)$, that arises from
${\mathcal B}_\varepsilon$ by $T$-duality on the fibers of the Hitchin
fibration.
\def{\mathcal H}{{\mathcal H}}
We will postpone a fuller explanation to section
\ref{operbrane}, but in brief the dual of ${\mathcal B}_\varepsilon$ is a
Lagrangian brane ${\mathcal B}_{{\cmmib N}}$ supported on a Lagrangian submanifold
${{\cmmib N}}$ that is known as the variety of opers. We can use this dual
description to describe the space ${\mathcal H}$ of $({\mathcal B}_\varepsilon,{\mathcal B}_L)$
strings; it is the same as the space of $({\mathcal B}_{{\cmmib N}},{\mathcal B}_{\widetilde L})$
strings. This space is easily described, assuming that the two
Lagrangian submanifolds ${{\cmmib N}}$ and $\widetilde L$ have generic
(transverse) intersections. ${\mathcal H}$ has a basis\footnote{In
general, in the $A$-model of a symplectic manifold $X$,
world-sheet instanton effects can remove from the cohomology the
states corresponding to some intersections. In the present
context, both branes ${\mathcal B}_{{\cmmib N}}$ and ${\mathcal B}_{\widetilde L}$ are of type
$(A,B,A)$, so they have extra supersymmetry in common beyond what
is typical in an $A$-model. This extra supersymmetry generates an
extra fermion zero mode in the field of an instanton, ensuring
that instanton effects do not alter the cohomology that one reads
off classically from the intersections of the two Lagrangian
submanifolds.} with one basis vector for every intersection point
of ${{\cmmib N}}$ and $\widetilde L$.
The commuting Hamiltonians of the integrable system correspond to
holomorphic functions on ${{\cmmib N}}$. The eigenvalues of the commuting
Hamiltonians are simply the values of the corresponding functions
at the points on ${{\cmmib N}}$ at which ${{\cmmib N}}$ intersects $\widetilde L$.
Differently put, the joint eigenstates of the commuting
Hamiltonians correspond to opers (points in ${{\cmmib N}}$) that obey a
certain system of equations stating that they lie in $\widetilde L$.
As remarked above, there is an example \cite{F1,FFR,F2,F3} of an
integrable system whose spectrum has been described in just such a
fashion.
\section{Conformal Blocks From Four Dimensions}\label{confblocks}
In this section, we will use similar methods to study a different
problem -- the relation \cite{AGT} of four-dimensional gauge
theory to Liouville theory and like theories in two dimensions.
\subsection{Gauge Theory And Liouville Theory}\label{gl}
We consider an ${\mathcal N}=2$ supersymmetric gauge theory in four
dimensions on the four-manifold $M'={\mathbb R}\times S^3$. Because there
are now three curved dimensions in spacetime (as opposed to two in
section \ref{compom}, where we worked on $M={\mathbb R}\times S^1\times
D_R$), topological twisting now leaves only two unbroken
supercharges. This is so for any product metric on ${\mathbb R}\times
S^3$, regardless of the choice of metric on $S^3$. One unbroken
supercharge is the usual supercharge $Q$ that is related to
Donaldson theory (in the case of $SU(2)$ gauge theory without
hypermultiplets). In Euclidean signature, the second supercharge
$\overline Q$ can be obtained from the first by an
orientation-reversing reflection of the first factor ${\mathbb R}$ of $M'$.
In Lorentz signature, $\overline Q$ is the hermitian adjoint of $Q$.
The twisting that preserves $Q$ and $\overline Q$ is defined in the
usual way, by identifying the $SU(2)_R$ global symmetry group with
the structure group of the spin bundle of $S^3$. If one modifies
the geometry of $M'={\mathbb R}\times S^3$ so as to be no longer a
product, then the positive and negative chirality spin bundles of
$M'$ become distinct. For twisting, we must then decide whether
we want to identify $SU(2)_R$ with the structure group of the
positive or negative chirality spin bundle. As a result, in the
twisted theory with a non-product metric, we can conserve either
$Q$ or $\overline Q$ but not both.
The most important special case of this arises if we ``cap off''
the metric on $M'$. Topologically, we do this by viewing $S^3$ as
the boundary of a ball $B^4$. Instead of a flat metric on $B^4$,
we pick a metric that looks near the boundary like a product
${\mathbb R}^-\times S^3$, where ${\mathbb R}^-$ is a half-line $t\leq 0$, but such
that $S^3$ shrinks to a point at some $t=t_0<0$. In defining a
topologically twisted theory on $B^4$ with such a metric, we can
preserve $Q$ or $\overline Q$ but not both.
The basic property of $S^3$ that we will use is that it admits a
$U(1)\times U(1)$ action. If $S^3$ is viewed as the locus
$\sum_{i=1}^4 y_i^2=1$ in ${\mathbb R}^4$, then we introduce polar
coordinates $y_1+iy_2=ue^{i\alpha}$, $y_3+iy_4=ve^{i\beta}$ and
finally $(u,v)=(\sin w,\cos w)$. $U(1)\times U(1)$ acts by shifts
of $\alpha$ and $\beta$. The round metric on $S^3$ is ${\mathrm d}
w^2+\sin^2 w \,{\mathrm d}\alpha^2+\cos^2 w\,{\mathrm d}\beta^2$. It will be
convenient for us to use a more general $U(1)\times
U(1)$-invariant metric on $S^3$. We let $w$ run over an interval
$0\leq w\leq \ell$, and we take
\begin{equation}\label{pokk}{\mathrm d} s_{S^3}^2={\mathrm d} w^2+ f(w) {\mathrm d}\alpha^2
+g(w){\mathrm d}\beta^2,\end{equation} where $f(w)=\rho_1^2$ except very
near the left end-point $w=0$ where it vanishes quadratically, and
$g(w)=\rho_2^2$ except very near the right end-point $w=\ell$
where it vanishes quadratically. Here $\rho_1$ and $\rho_2$ are
constants. In describing $S^3$ as in (\ref{pokk}), we are in
effect viewing it as a ``warped'' $T^2$ fibration over a
one-dimensional base. On $M'={\mathbb R}\times S^3$, we take the product
metric ${\mathrm d} s^2={\mathrm d} t^2+{\mathrm d} s_{S^3}^2$.
\subsection{$\Omega$-Deformation}\label{omegadef}
What we want to do with this theory is to $\Omega$-deform it,
using the $U(1)\times U(1)$ symmetry. We follow the same basic
procedure as in (\ref{frog}), but now we use the fact that
four-dimensional gauge theories with ${\mathcal N}=2$ supersymmetry can
arise by dimensional reduction from a six-dimensional theory with
two more coordinates that we will here call $x^4$ and $x^5$. Apart
from adding $({\mathrm d} x^4)^2+({\mathrm d} x^5)^2$ to the metric ${\mathrm d} s^2$, the
only change that we make is to modify the terms that involve ${\mathrm d}
\alpha$ and ${\mathrm d} \beta$. A simple case is a modification in which
${\mathrm d}\alpha$ ``mixes'' only with ${\mathrm d} x^4$ and ${\mathrm d}\beta$ ``mixes''
only with ${\mathrm d} x^5$, in the sense that the relevant part of the
six-dimensional metric is
\begin{equation}\label{gunn} f(w)\left({\mathrm d}\alpha -\varepsilon_1 {\mathrm d}
x^4)^2+g(w)\right({\mathrm d}\beta-\varepsilon_2 {\mathrm d} x^5)^2+({\mathrm d} x^4)^2+({\mathrm d} x^5)^2.\end{equation}
We have considered here a ``diagonal'' deformation in which
$\alpha$ and $\beta$ mix only with $x^4$ and $x^5$, respectively.
We take the deformation parameters $\varepsilon_1$ and
$\varepsilon_2$ to be real and positive. In greater generality, as
in \cite{Nek}, one can introduce complex deformation parameters
$\epsilon_1$, $\epsilon_2$, set $x=x^4+ix^5$ and generalize
(\ref{gunn}) to
\begin{equation}\label{punn}f(w)\left({\mathrm d}\alpha-{\mathrm{Re}}(\overline\epsilon_1
{\mathrm d} x)^2\right)+g(w)\left({\mathrm d}\beta-{\mathrm{Re}}(\overline\epsilon_2{\mathrm d}
x)^2\right)+({\mathrm d} x^4)^2+({\mathrm d} x^5)^2.\end{equation} Evidently, the special case
(\ref{gunn}) corresponds to \begin{equation}\label{potz}
\epsilon_1=\varepsilon_1,~~\epsilon_2=i\varepsilon_2,\end{equation}
and so
\begin{equation}\label{unn}\frac{\epsilon_2}{\epsilon_1}=i\frac{\varepsilon_2}{\varepsilon_1}.\end{equation}
We will analyze here the special case of a diagonal
deformation, and then in appendix \ref{gencompgeom}, we will
analyze the general case using generalized complex geometry.
\defS^3_{\varepsilon_1,\varepsilon_2}{S^3_{\varepsilon_1,\varepsilon_2}}
\def\H_{\varepsilon_1,\varepsilon_2}{{\mathcal H}_{\varepsilon_1,\varepsilon_2}}
\defB^4_{\varepsilon_1,\varepsilon_2}{B^4_{\varepsilon_1,\varepsilon_2}}
\defS^4_{\varepsilon_1,\varepsilon_2}{S^4_{\varepsilon_1,\varepsilon_2}}
\def\R^4_{\varepsilon_1,\varepsilon_2}{{\mathbb R}^4_{\varepsilon_1,\varepsilon_2}}
Let $S^3_{\varepsilon_1,\varepsilon_2}$ be $S^3$ with the
$\Omega$-deformation with the indicated parameters. We let
${\mathcal H}_{\varepsilon_1,\varepsilon_2}$ be the space of supersymmetric
ground states of the $\Omega$-deformed theory on ${\mathbb R}\times S^3_{\varepsilon_1,\varepsilon_2}$.
We interpret the results of \cite{AGT} to mean that $\H_{\varepsilon_1,\varepsilon_2}$ can be
identified, for $G=SU(2)$, with the space of Virasoro conformal
blocks on the Riemann surface $C$. (Virasoro conformal blocks are
the conformal blocks of Liouville theory. If $SU(2)$ is replaced
by another simply-laced group $G$, then \cite{Wyllard,MMMM} the
relevant objects are the conformal blocks of the corresponding
Toda field theory.) Understanding this claim and certain related
facts will be our goal in the rest of this paper.
Our first step will be to follow the logic of section \ref{alternative}
and compare the $\Omega$-deformed description to a standard one.
In the region in which $f$ and $g$ are
constants, the $\Omega$-deformation can be removed by an ordinary
change of variables. The combined operation of
$\Omega$-deformation plus ordinary change of variables amounts to
a rotation of the $\alpha 4$ plane times a rotation of the $\beta
5$ plane, each of them precisely analogous to (\ref{talign}). (The
more general deformation (\ref{punn}) can also be removed by a
change of variables in the region in which $f$ and $g$ are
constants, though the necessary formulas are more complicated.)
Just as in section \ref{alternative}, the simplest situation is
that in which the radii $\rho_1$, $\rho_2$ in the
$\Omega$-deformed description are taken to infinity, keeping the
deformation parameters $\varepsilon_i$ fixed. Then the rotation
that goes from $\Omega$-deformed description to the one by
conventional variables is simply the product of $\pi/2$ rotations
in the $\alpha4$ and $\beta5$ planes. In the undeformed language,
the radii of the circles parametrized by $\alpha$ and $\beta$ are
\begin{equation}\label{properly}\widehat\rho_1=\frac{1}{\varepsilon_1},~~\widehat\rho_2
=\frac{1}{\varepsilon_2}.\end{equation}
This means that the $\tau$ parameter of the torus parametrized by
$\alpha,\beta$ is in the undeformed description
\begin{equation}\label{roperly}\widehat\tau=i\frac{\varepsilon_2}{\varepsilon_1}
=\frac{\epsilon_2}{\epsilon_1}.\end{equation} (We denote this
$\tau$ parameter as $\widehat\tau$ to avoid confusion with generic
coupling parameters of an ${\mathcal N}=2$ supersymmetric gauge theory, or
complex structure parameters of a Riemann surface $C$, which we
have called $\tau$.)
Extending (\ref{nelf}), the gauge coupling in the ordinary
description is
\begin{equation}\label{troperly}\widehat
e^2=\frac{e^2}{\sqrt{(1+\varepsilon_1^2\rho_1^2)(1+\varepsilon_2^2\rho_2^2)}}.\end{equation}
In particular, $\widehat e^2\to 0$ as $\rho_1,\rho_2\to\infty$.
The $\pi/2$ rotations means that in transforming from the
$\Omega$-deformed description to the ordinary description, the
supersymmetry generator $\eta$ is transformed by
\begin{equation}\label{zoperly}\eta\to \frac{1+\Gamma_{\alpha 4}}{\sqrt 2}\frac{1+\Gamma_{\beta
5}}{\sqrt 2}\eta.\end{equation}
\subsection{Two-Dimensional Description}\label{zoth}
Just as in section \ref{compom}, we want to think of $M'={\mathbb R}\times
S^3$ as an $S^1\times \widetilde S^1$ fibration, with fiber
parametrized by $\alpha,\beta$, over ${\mathbb R}\times I$, parametrized by
$t,w$. As in section \ref{compom}, there are branes at both ends
of $I$. In section \ref{compom}, just one of these branes
originated from the geometry at the tip of a cigar. But in our
present problem, both branes have such an origin. In fact, near
either end of $I$, $M'$ resembles, up to a fairly obvious change
of variables, the cigar geometry that led to our friend
${\mathcal B}_\varepsilon$.
This will enable us to borrow the analysis of section
\ref{compom}, but we do have to keep in mind that the isomorphism
with section \ref{compom} is different at the two ends of $I$.
Near $w=0$, the $\alpha$ circle is shrinking and thus corresponds
to $\widetilde S^1$ in section \ref{compom}. The coordinates
$(t,w,\alpha,\beta)$ should be matched near $w=0$ with
$(x^0,x^2,x^3,x^1)$ in section \ref{compom}. Near $w=\ell$, the
$\beta$ circle shrinks and corresponds to $\widetilde S^1$; the roles
of $\alpha$ and $\beta$ are exchanged.
\def\varepsilon{\varepsilon}
\def{\alpha}{{\alpha}}
\def{\beta}{{\beta}}
We will denote as ${\mathcal B}_{\alpha}$ the brane at $w=0$ where the
$\alpha$ circle shrinks, and as ${\mathcal B}_{\beta}$ the brane at $w=\ell$
where the $\beta$ circle shrinks. Let us first understand the
brane ${\mathcal B}_\alpha$ in terms of ordinary variables.
Since the geometry near $w=0$ is the same as the geometry near
$x^2=0$ in section \ref{compom}, the only reason that the brane
${\mathcal B}_\alpha$ is not trivially equivalent to the brane
${\mathcal B}_\varepsilon$ studied in section \ref{compom} is that a
rotation of observables must be made to compare the two. For
${\mathcal B}_\alpha$, the rotation that we want to make from
$\Omega$-deformed to ordinary variables is by the product
\begin{equation}\label{omurx}\frac{1+\Gamma_{\alpha 4}}{\sqrt 2}\frac{1+\Gamma_{\beta 5}}{\sqrt 2},\end{equation}
which in the
notation of section \ref{compom} corresponds to
\begin{equation}\label{zomurx}\frac{1+\Gamma_{34}}{\sqrt 2}\frac{1+\Gamma_{15}}{\sqrt 2},\end{equation}
By contrast, in section \ref{compom}, we
made the rotation by $(1+\Gamma_{34})/\sqrt 2$. So in comparing
${\mathcal B}_\varepsilon$ to ${\mathcal B}_\alpha$, we need to make an additional
rotation of the supersymmetry parameter $\eta$ (and all other
operators and observables) by $(1+\Gamma_{15})/\sqrt 2$, whose
effect is to exchange $\Gamma_1$ and $\Gamma_5$.
With this in mind, we re-examine some of the key equations of
section \ref{deformed}. The supersymmetries preserved by
${\mathcal B}_\varepsilon$ were characterized by eqn. (\ref{zumok}), and as
the matrix on the left hand side commutes with $\Gamma_{15}$, the
brane
${\mathcal B}_\alpha$ preserves the same supersymmetries as
${\mathcal B}_\varepsilon$. Likewise, the condition (\ref{urft}) that
characterizes which supersymmetries of a given brane can be
interpreted in two-dimensional topological field theory is again
defined with a matrix that commutes with $\Gamma_{15}$. So
${\mathcal B}_\alpha$ is a brane of type $(A,B,A)$, just like
${\mathcal B}_\varepsilon$.
The only difference between the two branes is that the $15$
rotation acts non-trivially on the $\mathbb{CP}^1$ that parametrizes
the possible topological field theory structure. Eqn.
(\ref{plumkin}), which singles out the supersymmetry generator of
the $B$-model in complex structure $J$, is again invariant under
the 15 rotation. But the other two conditions (\ref{guro}) and
(\ref{puro}), which characterize the $A$-models of types
$\omega_K$ and $\omega_I$, are exchanged under this
rotation.\footnote{\label{urtz} To be more precise, there is a
minus sign here, and the mapping takes
$(\omega_I,\omega_K)\to(\omega_K,-\omega_I)$. Similarly, two
paragraphs below, the exchange of $J$ and $K$ is really $(J,K)\to
(K,-J)$.}
The important consequence of this is that the supersymmetry charge
$Q$ that is related to instanton counting and Donaldson theory,
and which in section \ref{compom} was interpreted in undeformed
variables as the generator of the $A$-model supersymmetry of type
$\omega_K$, will now have to be interpreted as the generator of
the $A$-model supersymmetry of type $\omega_I$. This is the
supercharge we care about for our present purposes, since it is
the only one (apart from its adjoint) that is conserved in the
twisted theory on the full ${\mathbb R}\times S^3$ geometry.
Now let us consider our other brane ${\mathcal B}_\beta$, the one at
$w=\ell$. All the same reasoning applies except that $x^4$ is
exchanged with $x^5$. This exchanges the scalar field $A_4$ with
$A_5$. That exchange can be carried out by a $U(1)_R$
transformation which also rotates complex structure $J$ into $K$
(with a sign mentioned in footnote \ref{urtz}) while fixing $I$.
So ${\mathcal B}_{\beta}$ is a brane of type $(A,A,B)$. (The $U(1)_R$
transformation that maps $A_4$ to $A_5$ may not be a symmetry, as
it also acts on mass parameters in an underlying four-dimensional
Lagrangian. But still it can be used to determine the type of
supersymmetry preserved by ${\mathcal B}_{\beta}$, given that we know this
for ${\mathcal B}_{\alpha}$.) The same reasoning as before shows that the
usual supercharge $Q$ of instanton counting and Donaldson theory
is associated to the $A$-model of type $\omega_I$.
Obviously, the only structure that a brane ${\mathcal B}_\alpha$ of type
$(A,B,A)$ and a brane ${\mathcal B}_\beta$ of type $(A,A,B)$ have in common
is the $A$-model of type $\omega_I$. Everything hangs together,
since the supersymmetry generator of this $A$-model is indeed the
supercharge $Q$ that is conserved in the ${\mathbb R}\times S^3$ geometry.
\subsection{The Kahler Parameter Of The Sigma Model}\label{zelf}
Of the three Kahler classes on ${\mathcal M}_H$, $\omega_I$ is topologically
non-trivial, but $\omega_J$ and $\omega_K$ are topologically
trivial. This has been explained in section \ref{distinguished}.
Accordingly, the $A$-models of type $\omega_J$ or $\omega_K$ have
no coupling parameter. But the $A$-model of type $\omega_I$ does
have such a parameter, which we can think of as the Kahler
parameter in complex structure $I$. According to a standard
result \cite{HM,BJV} that was explained in section \ref{morec},
this parameter, assuming its imaginary part is positive, is the
modular parameter $\widehat\tau$ of the two-torus $T^2$ that is
parametrized by the angles $\alpha,\beta$, measured in the metric
that is appropriate for the undeformed description.\footnote{As
was explained in section \ref{distinguished} and as we further
explain at the end of the present subsection, this Kahler
parameter is complex-valued, with no condition of positivity, but
we have to use a different description when its imaginary part is
not positive.} That modular parameter was computed in eqn.
(\ref{roperly}):
\begin{equation}\label{zongor}\widehat\tau=i\frac{\varepsilon_2}{\varepsilon_1}
=\frac{\epsilon_2}{\epsilon_1}.\end{equation} So far we have only
deduced that the model is the $A$-model of type $I$ with this
$\widehat\tau$ parameter for real $\varepsilon_1, \varepsilon_2$.
However, this must be true in general by holomorphy. This will be
shown more explicitly in appendix \ref{gencompgeom} using
generalized complex geometry.
In particular, the exchange of the $\alpha$ and $\beta$ circles
corresponds to $\widehat\tau\to 1/\widehat\tau$. One might expect that
there would be a minus sign in this formula, so let us examine
this point closely. A simple exchange
$\alpha\leftrightarrow\beta$ (with no minus signs) gives an
orientation-preserving automorphism of the $S^3$ metric
(\ref{pokk}) if accompanied by $w\to\ell-w$. A look at
(\ref{punn}) shows that to extend this operation to a symmetry of
the $\Omega$-deformed theory, we must take
$\epsilon_1\leftrightarrow\epsilon_2$, again with no minus signs.
To get a symmetry that preserves the overall orientation of
${\mathbb R}\times S^3_{{\alpha},{\beta}}$, we take it to act trivially on
${\mathbb R}$. We call the combined operation $\cmmib W$. It acts on
$\widehat\tau$ by $\widehat\tau\to +1/\widehat\tau$.
$\cmmib W$ has an interesting interpretation in the context of
generalized quiver gauge theories. Consider a four-dimensional
${\mathcal N}=2$ theory obtained as in \cite{DG} by compactifying the
six-dimensional $(0,2)$ theory on a Riemann surface $C$. Then
compactify further to two dimensions on the $\alpha\beta$ torus
$T^2$.
Altogether, we are compactifying from six to two dimensions on
$T^2\times C$. Introducing an interval $I$ parametrized by $w$,
the full spacetime, away from the boundaries of $I$, is ${\mathbb R}\times
I\times C\times T^2$. If we carry out first the compactification
on $T^2$, we get ${\mathcal N}=4$ super Yang-Mills theory on ${\mathbb R}\times
I\times C$. The orientation-preserving mapping class group of
$T^2$ is $SL(2,{\Bbb Z})$, and this is usually called the
electric-magnetic duality group in four dimensions. However, if
we allow diffeomorphisms of $T^2$ that reverse its orientation, we
get an extended mapping class group $GL(2,{\Bbb Z})$.
Orientation-reversing symmetries of $ T^2$ lead to symmetries of
${\mathcal N}=4$ super Yang-Mills theory if accompanied by
orientation-reversing symmetries of spacetime. In the present
case, $\cmmib W$ acts by $\alpha\leftrightarrow\beta$, which
reverses the orientation of $T^2$, and is accompanied by a
transformation $w\to\ell-w$ which reverses the orientation of
${\mathbb R}\times I\times C$.
We can relate what we have found to a more familiar symmetry
$\widehat\tau\to -1/\widehat\tau$. Let $\cmmib T$ be the combination
$(\alpha,\beta)\to(-\alpha,\beta)$ together with time-reversal
acting on ${\mathbb R}$ and the identity on $w$. Then $\cmmib T$ preserves
the orientation of ${\mathbb R}\times S^3_{{\alpha},{\beta}}$. It acts on
the deformation parameters by $(\epsilon_1,\epsilon_2)\to
(-\epsilon_1,\epsilon_2)$. So the combination $\cmmib T \cmmib W$
acts on the deformation parameters by
$(\epsilon_1,\epsilon_2)\to (\epsilon_2,-\epsilon_1)$, or
$\widehat\tau\to -1/\widehat\tau$.
In the description by ${\mathcal N}=4$ super Yang-Mills theory on ${\mathbb R}\times
I\times C$, $\cmmib T$ is time-reversal, acting as $-1$ on ${\mathbb R}$
and trivially on $I\times C$. The composition $\cmmib{TW}$
preserves the orientation of ${\mathbb R}\times I\times C$, and acts on
$\widehat\tau$ as a standard electric-magnetic duality transformation
$\widehat\tau\to-1/\widehat\tau$.
According to \cite{KW}, in the context of compactification from
four to two dimensions on a Riemann surface $C$, the symmetry
$\widehat\tau\to -1/\widehat\tau$ is the basis for geometric Langlands
duality. The most basic application of this duality to the
geometric Langlands program involves the duality between the
$B$-model of ${\mathcal M}_H$ in complex structure $J$ and the $A$-model of
type $\omega_K$. Section \ref{compom} of the present paper was
based on the $A$-model of ${\mathcal M}_H$ of type $\omega_K$, but its
duality with the $B$-model of type $J$ did not play an important
role.
However, geometric Langlands duality has an extension\footnote{In
the mathematical literature, this extension is often called
quantum geometric Langlands, but we will avoid this terminology
because we interpret also ``classical'' geometric Langlands via
quantum field theory.} involving a complex parameter that was
called $\Psi$ in \cite{KW}. As explained in section 5.2 of
\cite{KW}, extended geometric Langlands can be naturally described
by the $A$-model of ${\mathcal M}_H$ in symplectic structure $\omega_I$,
with the modulus $\widehat\tau$ of that $A$-model equal to $\Psi$. The
operation $\cmmib{TW}$ amounts to the extended geometric Langlands
duality $\Psi\to -1/\Psi$. Thus, the exchange of the two circles,
which certainly will be important in our analysis in the rest of
this paper, is the basis for extended geometric Langlands duality.
In the $A$-model of type $\omega_I$ as usually defined,
$\widehat\tau=\Psi$ is a Kahler parameter that takes values in the
upper half of the complex plane. However, according to \cite{KW},
and as we explained in relation to eqn. (\ref{oplo}), in a more
complete description, $\Psi$ actually parametrizes $\mathbb{CP}^1$.
Values of $\Psi$ in the lower half plane can be reached by an
$A$-model of symplectic structure $-\omega_I$. The cases that
$\Psi$ is real or $\infty$ are more subtle. The case of real
$\Psi$ can be studied as an $A$-model with symplectic form
$\omega_K$ and a $B$-field proportional to $\omega_I$, while
$\Psi=\infty$ is the $B$-model of type $J$.
\subsection{More About The Branes}\label{onko}
Now let us discuss in more detail the branes that appear in our
$T^2$ compactification to ${\mathbb R}\times I$.
To arrive at a conventional sigma-model on ${\mathbb R}\times I$, we must
make a $T$-duality on the scalars that arise from gauge field
holonomies on either the $\alpha$ circle or the $\beta$ circle.
Suppose that we make the $T$-duality associated to the $\alpha$
circle. Then as explained in section \ref{zoth}, the brane
${\mathcal B}_{\alpha}$ is a rotated version of the brane of type $(A,B,A)$
studied in section \ref{compom}.
If we make the opposite $T$-duality on the $\beta$ circle, the
other brane ${\mathcal B}_{\beta}$ can be described similarly; it is the
analogous space-filling rank 1 brane of type $(A,A,B)$. By a
$U(1)_R$ chiral rotation (which in general will transform the mass
and scale parameters of the theory), we can rotate $K$ back to
$J$ and arrive at a description just like the one in the last
paragraph. Thus, ${\mathcal B}_\beta$ in one description is equivalent to
${\mathcal B}_\alpha$ in the other description, up to the chiral rotation.
However, our goal is to understand the physical Hilbert space
$\H_{\varepsilon_1,\varepsilon_2}$ of an underlying four-dimensional ${\mathcal N}=2$ theory
compactified on $S^3$ with an $\Omega$-deformation. From a
two-dimensional viewpoint, ${\mathcal H}$ is the space of
$({\mathcal B}_\alpha,{\mathcal B}_\beta)$ strings. To describe this space, we need to
describe both branes in the same language.
We have two options. We can describe ${\mathcal B}_\alpha$ via a
$T$-duality associated to the $\beta$ circle rather than the
$\alpha$ circle. This will turn ${\mathcal B}_\alpha$ into a Lagrangian
brane ${\mathcal B}_{{{\cmmib N}}}$ of type $(A,B,A)$ that we will describe.
Or conversely we can describe ${\mathcal B}_\beta$ via a $T$-duality on the
$\alpha$ circle. In this description, ${\mathcal B}_\beta$ turns into a
Lagrangian brane ${\mathcal B}_{{{\cmmib N}}'}$ of type $(A,A,B)$ that can be reached
from ${\mathcal B}_{{\cmmib N}}$ by a $U(1)_R$ rotation.
The space ${\mathcal H}$ has two dual descriptions. It is the space of
$({\mathcal B}_{{\cmmib N}},{\mathcal B}_\beta)$ strings in the $A$-model of type $\omega_I$
with coupling parameter $\widehat\tau=i\varepsilon_1/\varepsilon_2$,
or the space of $({\mathcal B}_\alpha,{\mathcal B}_{{{\cmmib N}}'})$ strings in the ``same''
model but with $\widehat\tau=i\varepsilon_2/\varepsilon_1$.
We have put quotes around the word ``same'' because there is
actually a duality involved. The two descriptions differ by the
combined $T$-duality associated to the $\alpha$ and $\beta$
circles (which we recall is the basic geometric Langlands
duality); this is the same as a $T$-duality on the fibers of the
Seiberg-Witten fibration. In the simplest generalized quiver
theories that originate in six dimensions, the Seiberg-Witten
fibration is the Hitchin fibration for Higgs bundles with a
simply-laced gauge group $G$. Under duality on the fibers of the
Hitchin fibration, the structure group $G$ of the Hitchin
fibration is mapped to $G^\vee$. (This was first shown in
\cite{HT}.) Since $G$ is simply-laced, $G$ and $G^\vee$ have the
same universal cover; if one is $SU(2)$, the other is $SO(3)$.
Still, the distinction between $G$ and $G^\vee$ is significant in
a very precise description.
\begin{table}
\begin{center}
\begin{tabular}{c|c}
Model&Dual\\
\hline
${ I_B}$&${ I_B}$\\
${I_A}$&${ I_A}$\\ ${ J_B}$&${ K_A}$\\ ${
J_A}$&${ K_B}$\\ ${ K_B}$&${ J_A}$\\ ${
K_A}$&${ J_B}$\\
\end{tabular}
\end{center}
\begin{caption}\noindent\small{
Listed here are the $A$- and $B$-models to which a given model
transforms under $T$-duality on the fibers of the Hitchin
fibration. For example, the last row asserts that $K_A$, the
$A$-model of type $\omega_K$, is mapped to $J_B$, the $B$-model of
type $J$.}
\end{caption}
\end{table}
The combined $T$-duality on the $\alpha$ and $\beta$ circles maps
${\mathcal B}_\alpha$ to ${\mathcal B}_{{\cmmib N}}$ and ${\mathcal B}_\beta$ to ${\mathcal B}_{{{\cmmib N}}'}$. How it
acts on the supersymmetries is summarized in the table. For
example, the first row of the table asserts that the $B$-model of
type $I$ -- denoted in the table as $I_B$ -- is mapped to itself
by the $T$-duality on the fibers of the Hitchin fibration. This
reflects the fact that the Hitchin fibration is holomorphic in
complex structure $I$. Similarly, the $B$-model of type $J$ --
denoted $J_B$ -- is exchanged with the $A$-model of type
$\omega_K$ -- denoted $K_A$. The table is explained in section 5
of \cite{KW}.
\subsection{A Practice Case}\label{practice}
\def{v}{{v}}
Before discussing the duals of the branes ${\mathcal B}_\alpha$ and
${\mathcal B}_\beta$, we will first practice with a simpler but still subtle
case. This simpler case is the brane ${\mathcal B}_*$ of type $(B,B,B)$
whose support is all of ${\mathcal M}_H$ and whose Chan-Paton bundle is
trivial, by which we mean in particular that it is flat.
Since the Chan-Paton bundle of ${\mathcal B}_*$ is flat, it is certainly
flat when restricted to each fiber of the Hitchin fibration. So a
fiber of the Hitchin fibration is mapped by the duality to a
point, and hence the $T$-dual ${\mathcal B}_{{{\cmmib N}}^*}$ of ${\mathcal B}_*$ is supported
on a section ${{\cmmib N}}^*$ of the Hitchin fibration (of the dual group
$G^\vee$). ${{\cmmib N}}^*$ is automatically middle-dimensional, and as
${\mathcal B}_{{{\cmmib N}}^*}$ is a brane of type $(B,A,A)$, it is holomorphic in
complex structure $I$ and Lagrangian with respect to
$\Omega_I=\omega_J+i\omega_K$.
The section ${{\cmmib N}}^*$ of the Hitchin fibration is actually almost
uniquely determined (up to a choice in what follows of a spin
structure on $C$) by the fact that it is holomorphic in complex
structure $I$. How to construct such a holomorphic section was
shown in \cite{Hitchin}. We will describe the construction only
for $G=SU(2)$. We think of ${\mathcal M}_H$ as a moduli space of stable
Higgs bundles, that is pairs $(E,\varphi)$ where $E$ is a rank two
holomorphic bundle vector over $C$ of trivial determinant, and
$\varphi$ is a holomorphic section of $K_C\otimes
{\mathrm{ad}}(E)$. To describe a holomorphic section of the
Hitchin fibration, we pick a square root $K_C^{1/2}$, and take
$E=K_C^{-1/2}\oplus K_C^{1/2}$. This is the most unstable bundle
$E$ for which there exists a stable Higgs bundle $(E,\varphi)$.
Requiring $(E,\varphi)$ to be stable, the most general choice of
$\varphi$ up to an automorphism of $E$ is
\begin{equation}\label{telf}\varphi=\begin{pmatrix} 0 & 1\cr {v} & 0
\end{pmatrix},\end{equation} where ${v}$ is a quadratic differential on $C$ and we regard
$\varphi$ as a matrix acting on $\begin{pmatrix}K_C^{-1/2} \cr
K_C^{1/2}\end{pmatrix}$. For $G=SU(2)$, the fibers of the Hitchin
fibration are parametrized by the value of ${\mathrm {Tr}}\,\,\varphi^2$. Since
${\mathrm {Tr}}\,\,\varphi^2=2{v}$, there is one choice of ${v}$ for any
desired value of ${\mathrm {Tr}}\,\,\varphi^2$. Hence this family of Higgs
bundles $(E,\varphi)$, which we will call ${{\cmmib N}}^*$, gives a
section of the Hitchin fibration. This section is manifestly
holomorphic in complex structure $I$. To show that ${{\cmmib N}}^*$ is
Lagrangian for the holomorphic two-form $\Omega_I$, we use the
explicit formula
\begin{equation}\label{duzobox}\Omega_I=\delta\left(\frac{1}{\pi}\int_C{\mathrm {Tr}}\,\phi_z\delta
A_{\overline z}\right).\end{equation} Since $E$, which is characterized
by $A_{\overline z}$, is fixed in the family ${{\cmmib N}}^*$, we can take
$\delta A_{\overline z}=0$ when restricted to ${{\cmmib N}}^*$. So a brane
${\mathcal B}_{{{\cmmib N}}^*}$ supported on ${{\cmmib N}}^*$ with trivial Chan-Paton bundle
is a brane of type $(B,A,A)$, as expected.
Continuing with the case $G=SU(2)$, ${{\cmmib N}}^*$ can be naturally
identified with the Teichmuller space $C$ of the Riemann surface
$C$. (This is proved \cite{Hitchin}
by showing
that ${{\cmmib N}}^*$ can be interpreted as a component of the moduli space
of flat $SL(2,{\mathbb R})$ bundles on $C$.) In fact, the symplectic form
$\omega_I$ of ${\mathcal M}_H$, restricted to ${{\cmmib N}}^*$, and with the standard
normalization eqn. (\ref{obox}), is the natural Weil-Petersson
symplectic form of Teichmuller space. Since the coupling
parameter of our $A$-model is
$\widehat\tau=i\varepsilon_1/\varepsilon_2$, the symplectic form of
the $A$-model is $\varepsilon_1/\varepsilon_2$ times the
Weil-Petersson form.
The relation of $\omega_I$ to the Weil-Petersson form makes
possible the following construction, described in section 4 of
\cite{GW}. Let ${\mathcal B}'$ be the brane with support all of ${\mathcal M}_H$ and
Chan-Paton curvature
$\omega_I^*=(\varepsilon_1/\varepsilon_2)\omega_I$. So ${\mathcal B}'$ is a
coisotropic brane of type $(B,A,A)$.
Viewing ${\mathcal B}'$ and ${\mathcal B}_{{{\cmmib N}}^*}$ as $A$-branes of type $\omega_K$,
let ${\mathcal H}'$ be the space of $({\mathcal B}',{\mathcal B}_{{{\cmmib N}}^*})$ strings. As reviewed
in section \ref{coiso}, ${\mathcal H}'$ can be interpreted as quantization
of ${{\cmmib N}}^*$ with symplectic structure $\omega_I^*$ -- in other
words, quantization of Teichmuller space with symplectic form
$\varepsilon_1/\varepsilon_2$ times the Weil-Petersson form.
Now let us think back to the problem of four-dimensional gauge
theory that has motivated our analysis in this section: how to
relate the space $\H_{\varepsilon_1,\varepsilon_2}$ of supersymmetric ground states of a rank
1 generalized quiver theory on ${\mathbb R}\timesS^3_{\varepsilon_1,\varepsilon_2}$ to the space of
Virasoro (or Liouville) conformal blocks on $C$. Teichmuller
space has been quantized \cite{CF} using a real polarization
(which depends on the choice of a set of $A$-cycles on $C$) in a
way that certainly appears to give a good candidate for the space
of Virasoro conformal blocks. (There is also an analog for groups
of higher rank \cite{FG}.) If therefore we are aiming to get the
space of Virasoro conformal blocks as the space of ground states
of open strings stretched between two branes, then ${\mathcal B}'$ and
${\mathcal B}_{{{\cmmib N}}^*}$ would appear to be good candidates for those branes.
These are, however, not the candidates that have emerged from our
analysis. We have found instead that $\H_{\varepsilon_1,\varepsilon_2}$ is the space of
$({\mathcal B}_\alpha,{\mathcal B}_{{{\cmmib N}}'})$ strings (or the space of
$({\mathcal B}_{{{\cmmib N}}},{\mathcal B}_\beta)$ strings). Here ${\mathcal B}_\alpha$ is related to the
more naive candidate ${\mathcal B}'$ by a sort of hyper-Kahler rotation;
one is a rank one coisotropic brane of type $(A,B,A)$, and the
other is a similar object of type $(B,A,A)$. Similarly,
${\mathcal B}_{{{\cmmib N}}'}$ is related to ${\mathcal B}_{{{\cmmib N}}^*}$ by a sort of dual
hyper-Kahler rotation. We do not have an intuitive understanding
of why our construction has led to the space of
$({\mathcal B}_\alpha,{\mathcal B}_{{{\cmmib N}}'})$ strings rather than the more obvious
space of $({\mathcal B}',{\mathcal B}_{{{\cmmib N}}^*})$ strings. Perhaps these spaces are
actually naturally isomorphic.
The space of $({\mathcal B}',{\mathcal B}_{{{\cmmib N}}^*})$ strings, since it describes
quantization of a component of the moduli space of flat $SL(2,{\mathbb R})$
connections on the Riemann surface $C$, certainly appears to be
related to $SL(2,{\mathbb R})$ Chern-Simons gauge theory in three
dimensions. The construction also has an analog \cite{GW} that is
similarly related to $SU(2)$ Chern-Simons gauge theory in $2+1$
dimensions. In this analog, ${{\cmmib N}}^*$ is replaced by the locus of
flat $SU(2)$ bundles. This locus is characterized by the condition
$\varphi=0$ and is, like ${{\cmmib N}}^*$, the support of a Lagrangian
brane of type $(B,A,A)$.
\subsection{The Brane Of Opers}\label{operbrane}
Now we want to discuss the dual of ${\mathcal B}_\alpha$ (or equivalently,
modulo a $U(1)_R$ rotation, the dual of ${\mathcal B}_\beta$). The
Chan-Paton curvature of ${\mathcal B}_\alpha$ is not zero; rather it equals
$\omega_J^*=(\varepsilon_1/\varepsilon_2)\mathrm{Re}\,\Omega_I$.
However, as the fibers of the Hitchin fibration are Lagrangian for
$\Omega_I$, the Chan-Paton bundle of ${\mathcal B}_\alpha$ is flat when
restricted to a fiber of that fibration.
This means that when restricted to a fiber, the $T$-dual of
${\mathcal B}_\alpha$ is a point. Therefore, the $T$-dual of ${\mathcal B}_\alpha$ is
supported on a section of the Hitchin fibration. We call this
section $ {{\cmmib N}}$ and we write ${\mathcal B}_{{\cmmib N}}$ for the $T$-dual of
${\mathcal B}_\alpha$.
${\mathcal B}_{{\cmmib N}}$ is a brane of type $(A,B,A)$, and as its support is
middle-dimensional, it is a Lagrangian brane. Hence the
Chan-Paton bundle of ${\mathcal B}_{{\cmmib N}}$ is flat. It therefore is actually
trivial, since ${{\cmmib N}}$, being a section of the Hitchin fibration, is
equivalent topologically to the base $\cmmib B$ and so is
contractible.
\def{\mathcal F}{{\mathcal F}}
${{\cmmib N}}$ is different from the section ${{\cmmib N}}^*$ of the Hitchin
fibration that was described in eqn. (\ref{telf}), since they are
holomorphic in different complex structures. ${{\cmmib N}}^*$ is
holomorphic in complex structure $I$ but ${{\cmmib N}}$ is holomorphic in
complex structure $J$. The Hitchin fibration is holomorphic only
in complex structure $I$, and the fact that this is the complex
structure in which ${{\cmmib N}}^*$ is holomorphic makes ${{\cmmib N}}^*$ much
simpler to study than ${{\cmmib N}}$.
Nevertheless, an explicit description of ${{\cmmib N}}$ is known, in
essence, from work on geometric Langlands \cite{BD}. (In
addition, the dual of ${\mathcal B}_\alpha$ has been described in
four-dimensional gauge theory language in section 4 of \cite{GW2}.
It should be possible to reduce this description to two dimensions
and recover the result of \cite{BD}, though this has not yet been
done.) Since we do not know a quick path to this description and
will not use the details in the rest of this paper, we will here
simply state the result.
We describe the result only for gauge group $G=SU(2)$. Also we
assume that $C$ has a negative Euler class (if its genus is 0 or
1, we assume it has at least 3 or 1 marked points, respectively).
If $E\to C$ is a flat bundle of rank 2, it can in particular be
regarded as a holomorphic bundle. A flat bundle is called an
``oper'' if, viewed as a holomorphic bundle, it is a non-trivial
extension
\begin{equation}\label{polyg}0\to K_C^{1/2}\to E\to K_C^{-1/2}\to 0,\end{equation}
where $K_C^{1/2}$ is a square root of the canonical bundle of $C$.
For a given choice of $K_C^{1/2}$, a non-trivial extension of this
kind is unique up to isomorphism. If $E$ is a rank 2 bundle with
flat connection ${\mathcal A}$ that is an oper, then the $(0,1)$ part of
${\mathcal A}$ has the form
\begin{equation}\label{poldo}{\mathcal A}_{\overline z}=\begin{pmatrix}-a_{\overline z} & 0 \cr u & a_{\overline z}\end{pmatrix},\end{equation}
where $a_{\overline z}$ defines the complex structure of the line
bundle $K_C^{1/2}$, and $u$ is a $K$-valued $(0,1)$-form; the
choice of $u$ does not matter, up to gauge transformation, as long
as its cohomology class in $H^1(C,K_C))\cong{\Bbb C}$ is non-zero. It is
true, though not trivial, that for this ${\mathcal A}_{\overline z}$, it is
possible to pick ${\mathcal A}_z$ so that the curvature ${\mathcal F}_{z\overline
z}=\partial_z{\mathcal A}_{\overline z}-\partial_{\overline z}{\mathcal A}_z+[{\mathcal A}_z,{\mathcal A}_{\overline z}]$
vanishes. One simple fact is that if we write
\begin{equation}\label{mayo}{\mathcal A}_z=\begin{pmatrix} f& e\cr g &
-f\end{pmatrix},\end{equation} then the upper right matrix element
$e$ is a holomorphic function on $C$, which globally must be
constant. Looking at the diagonal part of the equation $F_{z\overline
z}=0$, we learn that $e$ must be nonzero.
Another simple fact is that given any choice of ${\mathcal A}_z$ that makes
the curvature vanish, any other choice can be obtained by the
shift
\begin{equation}\label{nolyp}{\mathcal A}_z\to {\mathcal A}_z+\begin{pmatrix}0 & 0 \\ w & 0 \end{pmatrix},\end{equation}
where $w$ is a quadratic differential.
So any two
opers differ by a quadratic differential. There is actually a
canonical way to map the space of opers to the space of quadratic
differentials, since the flat $SL(2,{\mathbb R})$ bundle that comes from
uniformization (that is, from the existence on $C$ of an Einstein
metric of constant negative curvature) can be interpreted as an
oper and gives a natural base point in the space of opers.
So for $G=SU(2)$, the space ${{\cmmib N}}$ of opers is, as a complex
manifold, naturally isomorphic to the space $H^0(C,K_C^2)$ of
quadratic differentials on $C$. (If there are marked points on
$C$, a similar derivation leads to quadratic differentials that
may have a pole of a specified type at the marked point.) For
generalized quiver theories \cite{DG} associated to $SU(2)$,
$H^0(C,K_C^2)$ is the same as the base of the Hitchin fibration,
which as usual we call $\cmmib B$. In general, for quiver gauge
theories based on any $G$, ${{\cmmib N}}$ is naturally isomorphic to
$\cmmib B$. (We should point out that there is something strange
about this assertion. The natural complex structure on ${{\cmmib N}}$ is
obtained by restricting to ${{\cmmib N}}$ the complex structure $J$ on
${\mathcal M}_H$, while the natural complex structure on $\cmmib B$ is
similarly related to $I$. Nevertheless, ${{\cmmib N}}$ with its natural
complex structure is naturally isomorphic to $\cmmib B$ with its
natural complex structure.)
Though we will not really use this information in the present
paper, the reader may find it helpful if we describe how opers are
related to conformal field theory. Again, we consider only the
case of $SU(2)$. First of all, locally it is possible to find a
gauge transformation of lower triangular form $\begin{pmatrix}1 &
0 \\ * & 1
\end{pmatrix}$ setting $u$ to zero and otherwise leaving ${\mathcal A}_{\overline
z}$ in the form (\ref{poldo}). In such a gauge ${\mathcal A}_z$ still has
the form (\ref{mayo}) and $e$ is still a nonzero constant; $f$ and
$g$ are now holomorphic sections of $K_C$ and $K_C^2$,
respectively. Without changing the form of ${\mathcal A}_{\overline z}$, we can
make a gauge transformation by a $2\times 2$ unimodular and
holomorphic matrix to set $e=1$ and $f=0$, whence
\begin{equation}\label{exoc} {\mathcal A}_z=\begin{pmatrix}0 & 1 \\ T &
0\end{pmatrix}\end{equation} where $ T$ is still holomorphic.
One might think that $T$ would be a
quadratic differential, but actually it is more naturally
understood as a stress tensor or projective connection. To see
why, consider an infinitesimal gauge transformation generated by
\begin{equation}\label{xoc}\begin{pmatrix} {\partial_z v}/2
& v\\ vT-\partial_z^2 v/2 & -\partial_z
v/2\end{pmatrix},\end{equation} with $v$ a holomorphic vector
field. A short calculation shows that this leaves fixed the form
of ${\mathcal A}_z$, and that $T$ transforms by
\begin{equation}\label{doc} T\to T+v\partial_z T +2(\partial_z v) T
-\frac{1}{2}\partial_z^3 v,\end{equation} in other words as a stress tensor.
\subsection{Physical States From The Brane Of
Opers}\label{physstate}
What we have described so far is the brane ${\mathcal B}_{{{\cmmib N}}}$ that is dual
to ${\mathcal B}_\alpha$. To find the dual to ${\mathcal B}_\beta$, which we call
${\mathcal B}_{{{\cmmib N}}'}$, we simply make a $U(1)_R$ rotation. Since
${\mathcal B}_{{{\cmmib N}}}$ is a Lagrangian brane of type $(A,B,A)$, the $U(1)_R$
rotation maps ${\mathcal B}_{{{\cmmib N}}}$ to a Lagrangian brane ${\mathcal B}_{{{\cmmib N}}'}$ of
type $(A,A,B)$, whose support ${{\cmmib N}}'$ is obtained from ${{\cmmib N}}$ by the
$U(1)_R$ rotation. Thus ${{\cmmib N}}'$ is a rotated version of the brane
of opers.
Our construction gives two related ways to describe the space
$\H_{\varepsilon_1,\varepsilon_2}$ of ground states of the ${\mathcal N}=2$ gauge theory on ${\mathbb R}\times
S^3_{\varepsilon_1,\varepsilon_2}$. It is the space of $({\mathcal B}_\alpha,{\mathcal B}_{{{\cmmib N}}'})$ strings (in a
description with gauge group $G$, $T$-duality on scalars related
to the $\alpha$ circle, and coupling parameter
$\varepsilon_1/\varepsilon_2$) or the space of $({\mathcal B}_{\widetilde
L},{\mathcal B}_\beta)$ strings (in a description with gauge group $G^\vee$,
$T$-duality on scalars related to the $\beta$ circle, and coupling
parameter $\varepsilon_2/\varepsilon_1$).
Either way, this sounds much like what we studied in section
\ref{compom}: one brane is a canonical coisotropic $A$-brane, and
the other is a Lagrangian $A$-brane. But there is a crucial
difference. Let us consider first the first description. Here
${\mathcal B}_\alpha$ is a coisotropic brane of type $(A,B,A)$ with
Chan-Paton curvature $F=\omega_J$, and ${\mathcal B}_{{{\cmmib N}}'}$ is a
Lagrangian brane of type $(A,A,B)$. Viewing these as $A$-branes
of type $\omega_I$, we want to describe the space $\H_{\varepsilon_1,\varepsilon_2}$ of
$({\mathcal B}_\alpha,{\mathcal B}_{{{\cmmib N}}'})$ strings.
If the Chan-Paton curvature $F$ were nondegenerate when restricted
to ${{\cmmib N}}'$, then $\H_{\varepsilon_1,\varepsilon_2}$ would arise by quantization of ${{\cmmib N}}'$ in
symplectic structure $F$; this was reviewed in section
\ref{coiso}. Here we are in the opposite situation. Since
$F=\omega_J$ and ${{\cmmib N}}'$ is Lagrangian for $\omega_J$, $F$
actually vanishes when restricted to ${{\cmmib N}}'$. This situation
sounds very special, but actually it is the usual situation
considered in geometric Langlands.
\def{\mathcal I}{{\mathcal I}}
In such a case, rather than by quantization, $\H_{\varepsilon_1,\varepsilon_2}$ can be
described as follows. Let ${\mathcal I}=\omega^{-1}F$ be the complex
structure determined by the coisotropic brane ${\mathcal B}_\alpha$. In the
present case, $\omega=\omega_I$, $F=\omega_J$, and ${\mathcal I}=K$. Note
that ${{\cmmib N}}'$ is a complex submanifold in complex structure $K$.
Then, roughly speaking, the space of physical states is the space
of holomorphic functions on ${{\cmmib N}}'$ (actually holomorphic sections
of a certain line bundle, as we explain momentarily), in complex
structure $K$. Since ${{\cmmib N}}'$ is as a complex manifold the same as
the base $\cmmib B$ of the Hitchin fibration, we can think of the
space $\H_{\varepsilon_1,\varepsilon_2}$ of physical states as the space of holomorphic
sections of a certain line bundle over $\cmmib B$.
To describe the relevant line bundle, observe that because of the
relation of $D$-branes to $K$-theory, by comparing the Chan-Paton
bundles of the two branes in question, one can extract a square
root $K^{1/2}_{{{\cmmib N}}'}$ of the canonical bundle of ${{\cmmib N}}'$.
$K^{1/2}_{{{\cmmib N}}'}$ is a holomorphic line bundle over ${{\cmmib N}}'$ that is
trivial, but not canonically so. $\H_{\varepsilon_1,\varepsilon_2}$ is the space of
holomorphic sections of $K^{1/2}_{{{\cmmib N}}'}$.
To justify this answer, we simply quantize open strings that
stretch between the two branes ${\mathcal B}_\alpha$ and ${\mathcal B}_{{{\cmmib N}}'}$. Let
the string worldsheet be ${\mathbb R}\times I$ where ${\mathbb R}$ is parametrized
by the time $t$, and $I$ is an interval, with boundary conditions
at the two ends set by the two branes. To find zero energy
states, we quantize the motion in time of the modes that have zero
kinetic energy along $I$. We can repeat the derivation in section
2.3 of \cite{GW}, but the result is now different because $F$
vanishes when restricted to ${{\cmmib N}}'$, rather than being
nondegenerate. The bosonic zero modes describe maps
$x:{\mathbb R}\to{{\cmmib N}}'$. There is no term first order in ${\mathrm d} x/{\mathrm d} t$,
because $F|_{{{\cmmib N}}'}=0$. So the low energy action for $x$ is the
usual sort of kinetic energy $\frac{1}{2}\int {\mathrm d} t g_{IJ}\frac{{\mathrm d}
x^I}{{\mathrm d} t}\frac{{\mathrm d} x^J}{{\mathrm d} t}$, where $g_{IJ}$ is the induced
metric on ${{\cmmib N}}'$. There also are fermionic zero modes $\psi^I$,
forming a section of the pullback by $x$ of the tangent bundle of
${{\cmmib N}}'$. To find this, one can follow eqns. (2.8)-(2.11) of
\cite{GW}, with the difference that now ${{\cmmib N}}'$ is real with
respect to $J$, rather than holomorphic as assumed in \cite{GW}.
While a key point in \cite{GW} was that there were no fermionic
zero modes, now zero modes $\psi^I$ do survive. The effective
action for $x^I,\psi^J$ must have two supercharges, descending
from the unbroken supercharges $Q$, $\overline Q$ of the gauge theory.
The minimal supersymmetric action for these fields with this amount of
supersymmetry is familiar:
\begin{equation}\label{zam}I=\int{\mathrm d} t\left(\frac{1}{2}g_{IJ}\frac{{\mathrm d}
x^I}{{\mathrm d} t}\frac{{\mathrm d} x^J}{{\mathrm d} t}+ig_{IJ}\psi^I\frac{D \psi^J}{D
t}\right).\end{equation} This is the basic sigma-model action in
one dimension, with two supercharges when (as here) the target
space is a Kahler manifold. In quantization, $Q$ becomes the
$\overline\partial$ operator acting on $(0,q)$-forms with values in
$K_{{{\cmmib N}}'}^{1/2}$, and $\overline Q$ is its adjoint. (The sum $\overline Q+
Q$ is the Dirac operator.) The cohomology of $Q$ is thus the
$\overline\partial$ cohomology of ${{\cmmib N}}'$ with values in
$K_{{{\cmmib N}}'}^{1/2}$. The cohomology of degree 0 consists of
holomorphic sections of $K_{{{\cmmib N}}'}^{1/2}$, and the higher
cohomology vanishes.
So finally, upon identifying ${{\cmmib N}}'$ as a complex manifold with
$\cmmib B$, the base of the Hitchin fibration, the space $\H_{\varepsilon_1,\varepsilon_2}$
of physical states can be identified as the space of holomorphic
sections of $K_{\cmmib B}^{1/2}$ over $\cmmib B$:
\begin{equation}\label{otung}\H_{\varepsilon_1,\varepsilon_2}=H^0({\cmmib B},K^{1/2}_{\cmmib
B}).\end{equation} There is a slight surprise here: the right hand
side does not depend on the parameters
$\varepsilon_1,\varepsilon_2$. However, the action of
observables on $\H_{\varepsilon_1,\varepsilon_2}$ does depend on these parameters, as we will
see shortly.
We obtained this result starting from the interpretation of
$\H_{\varepsilon_1,\varepsilon_2}$ as the space of $({\mathcal B}_\alpha,{\mathcal B}_{{{\cmmib N}}'})$ strings. The
alternative realization of $\H_{\varepsilon_1,\varepsilon_2}$ as the space of
$({\mathcal B}_{{{\cmmib N}}},{\mathcal B}_\beta)$ strings would lead after analogous steps to
a different identification of $\H_{\varepsilon_1,\varepsilon_2}$ with $H^0({\cmmib
B},K^{1/2}_{\cmmib B})$.
As in somewhat similar situations considered in section 2.3 of
\cite{GW}, the gauge theory on ${\mathbb R}\times S^3_{{\alpha},{\beta}}$,
or the two-dimensional sigma-model on ${\mathbb R}\times I$ to which it
reduces, will generate a hermitian inner product on
$\H_{\varepsilon_1,\varepsilon_2}=H^0({\cmmib B},K^{1/2}_{\cmmib B})$ making this space a
Hilbert space. But, except in the limit of large
$\varepsilon_1/\varepsilon_2$, when the Kahler modulus of ${\mathcal M}_H$
becomes large and the two-dimensional $\sigma$-model can be
treated semiclassically, this hermitian inner product is not
necessarily given by any elementary classical formula.
\subsection{Observables}\label{oddobservables}
Now we want to describe the algebra of observables that acts on
the space $\H_{\varepsilon_1,\varepsilon_2}$ of ground states of the gauge theory on $S^3_{\varepsilon_1,\varepsilon_2}$.
In the brane description, $\H_{\varepsilon_1,\varepsilon_2}$ is the space of
$({\mathcal B}_\alpha,{\mathcal B}_{{{\cmmib N}}})$ strings or alternatively the space of
$({\mathcal B}_{{{\cmmib N}}'},{\mathcal B}_\beta)$ strings. The first description makes it
clear that the algebra of $({\mathcal B}_\alpha,{\mathcal B}_\alpha)$ strings acts on
$\H_{\varepsilon_1,\varepsilon_2}$ by attaching to the left end of a string. The second
description makes it equally clear that the algebra of
$({\mathcal B}_\beta,{\mathcal B}_\beta)$ strings acts on $\H_{\varepsilon_1,\varepsilon_2}$ by attaching to the
right end of a string. The strings in question are the physical
states in the $A$-model of type $\omega_I$.
The two algebras of $({\mathcal B}_\alpha,{\mathcal B}_\alpha)$ strings and
$({\mathcal B}_\beta,{\mathcal B}_\beta)$ strings are equivalent, up to the usual
steps (a $U(1)_R$ rotation that may act on mass parameters, an
exchange $\varepsilon_1\leftrightarrow\varepsilon_2$, and duality
$G\leftrightarrow G^\vee$). So let us just describe the algebra
of $({\mathcal B}_\beta,{\mathcal B}_\beta)$ strings.
The brane ${\mathcal B}_\beta$ has Chan-Paton curvature
$F=(\varepsilon_1/\varepsilon_2)\omega_K$, and we study it in the
$A$-model of symplectic structure
$\omega=(\varepsilon_1/\varepsilon_2)\omega_I$. The
$({\mathcal B}_\beta,{\mathcal B}_\beta)$ strings are obtained by deformation
quantization of the ring ${\mathcal R}$ of holomorphic functions in complex
structure ${\mathcal I}=\omega^{-1}F=J$. To proceed farther, we focus on the
case of a generalized quiver theory associated to a Riemann
surface $C$. The two-dimensional description involves a
sigma-model with target ${\mathcal M}_H$, and in complex structure $J$,
${\mathcal M}_H$ is the moduli space of $\mathfrak g_{\Bbb C}$-valued flat connections
${\mathcal A}$ on $C$. The ring of holomorphic functions on ${\mathcal M}_H$ is
generated by traces of holonomies.\footnote{This fact is not
obvious but is proved for classical groups in \cite{AMR}, with the
understanding that by ``holomorphic functions'' on ${\mathcal M}_H$, we
really mean algebraic functions on ${\mathcal M}_H$, viewed as a moduli
space of representations of the fundamental group of the Riemann
surface $C$. Despite the fact that traces of holonomies suffice
to generate the ring, it may be more natural to consider also the
functions associated to labeled graphs, as for example in
\cite{Witten3}.} In other words, for $\gamma\subset C$ a simple
closed curve and $R$ a representation of $G$, we consider the
function
\begin{equation}\label{lovely}
W_R(\gamma)={\mathrm {Tr}}\,_R\,P\exp\left(-\oint_\gamma{\mathcal A}\right);\end{equation}
such
functions generate ${\mathcal R}$.
The deformation of the commutative ring ${\mathcal R}$ to a noncommutative
but still associative ring depends on one complex parameter,
usually called $q$. This deformation, which is constructed in
\cite{AMR2} (of course there are also many related constructions
from different points of view), appears in two familiar quantum
field theory problems. One is the problem\cite{Witten2} of
quantizing Chern-Simons gauge theory with compact gauge group $G$
on ${\mathbb R}\times C$, where $C$ is a Riemann surface. The phase space
is ${\mathcal M}$, the moduli space of flat unitary (that is $\mathfrak
g$-valued) connections $A$ on $C$. The basic operators on the
physical Hilbert space are Wilson loop operators. These are the
same traces of holonomies considered in the last paragraph, except
that they are functions on ${\mathcal M}$, the space of flat $\mathfrak
g$-valued connections, rather than its complexification ${\mathcal M}_H$.
However, traces of holonomies regarded as functions on ${\mathcal M}$ have
natural analytic continuations to holomorphic functions on ${\mathcal M}_H$.
(The analytic continuations are given by the same traces, now
evaluated for flat connections that may be complex-valued.) So the
ring of classical observables of Chern-Simons theory with compact
gauge group $G$ is actually the same as the ring ${\mathcal R}$ of complex
holonomies considered in the last paragraph. In quantizing
Chern-Simons gauge theory with gauge group $G$, ${\mathcal R}$ is deformed
to a noncommutative, associative algebra of quantum Wilson loop
operators. The deformation parameter is the Chern-Simons level
$k$, in terms of which one defines $q=\exp(2\pi i/(k+h))$, where
$h$ is the dual Coxeter number of $G$. A physical Hilbert space
on which the quantum algebra can act exists only for positive
integer $k$. But the deformed algebra ${\mathcal R}_q$ of Wilson loop
operators can be constructed as a function of a complex variable
$q$. The procedure involved can be understood as deformation
quantization (for example, see \cite{Andersen} for this
interpretation). But the situation is much better than a typical
example of deformation quantization: this is a favorable case in
which deformation quantization gives a deformation parametrized by
a complex variable $q$, not just a deformation over a formal power
series ring.
The second related problem in which one encounters the deformed
ring ${\mathcal R}_q$ of holonomy functions is two-dimensional conformal
field theory. The most basic case is current algebra of a compact
group $G$. Here a monodromy operation on the space of conformal
blocks was introduced in \cite{EVerlinde} (and this influenced
subsequent work on Chern-Simons theory). The monodromy operation
is defined by transporting a primary field in a representation $R$
of $G$ around a loop $\gamma\subset C$. In the correspondence
between two-dimensional conformal field theory and
three-dimensional Chern-Simons theory, the space of conformal
blocks maps to the physical Hilbert space of Chern-Simons theory,
and the conformal field theory monodromies map to the action of
Wilson loop operators of gauge theory.
The conformal field theory operation just mentioned has been
adapted to Liouville theory in \cite{AGGTV,DGOT}. The deformation
parameter is the Liouville coupling $b$, which corresponds to the
Chern-Simons level $k$ if Liouville theory is related to
$SL(2,{\mathbb R})$ current algebra. So the same associative algebra
${\mathcal R}_q$ of quantized holonomies that can be extracted from $SU(2)$
Chern-Simons theory acts on the conformal blocks of Liouville
theory. This algebra is also seen in quantization of Teichmuller
space. For gauge groups of higher rank, there is a similar
relation between the quantum-deformed algebra of holonomies in
Chern-Simons gauge theory with compact gauge group, and the
deformed algebra of holonomies that acts in the quantization of
the higher rank analogs of Teichmuller space \cite{CF,FG}.
There is, however, a fundamental difference between the case of a
compact symmetry group such as $SU(2)$ and the case of a
noncompact group such as $SL(2,{\mathbb R})$. In the case of a compact
symmetry group or gauge group $G$, the deformed algebra ${\mathcal R}_q$ of
holonomies acts irreducibly on the space of conformal blocks of
two-dimensional conformal field theory, or equivalently the
Hilbert space of Chern-Simons theory. For Liouville theory or
$SL(2,{\mathbb R})$ Chern-Simons theory, this is far from being true.
Instead, in its action on the space of Virasoro conformal blocks,
the algebra ${\mathcal R}_q$ commutes with a dual algebra ${\mathcal R}_{q'}$.
${\mathcal R}_{q'}$ is a second deformed algebra of holonomy operators,
with its parameter differing by the Liouville duality
$b\leftrightarrow b^{-1}$.
In our presentation here, the action of the second commuting
algebra is manifest, since the $({\mathcal B}_\alpha,{\mathcal B}_\alpha)$ strings
acting at one end commute with the $({\mathcal B}_\beta,{\mathcal B}_\beta)$ strings
acting at the other end.
The space of Virasoro conformal blocks can be characterized as an
irreducible module for two algebras ${\mathcal R}_q$ and ${\mathcal R}_{q'}$ of
$SL(2,{\mathbb R})$ holonomies at dual values of the parameters. The two
algebras are noncommutative but associative, and commute with each
other. The fact that we have found this structure strongly supports the idea that
the Hilbert space $\H_{\varepsilon_1,\varepsilon_2}$ of the generalized quiver theory on
$S^3_{\varepsilon_1,\varepsilon_2}$ is indeed the space of Virasoro (or Liouville) conformal
blocks, as first argued in \cite{AGT}.
\subsubsection{Wilson And 't Hooft Operators}\label{wth}
Concretely, the algebras ${\mathcal R}_q$ and ${\mathcal R}_{q'}$ are generated by
$Q$-invariant Wilson and 't Hooft operators that act at one end of
$S^3_{{\alpha},{\beta}}$ or the other. Let us return to the
explicit description (\ref{pokk}) of $S^3_{{\alpha},{\beta}}$. When
we toroidally reduce $S^3_{{\alpha},{\beta}}$ to an interval $I$
parametrized by $w$ with $0\leq w\leq\ell$, the end at $w=0$ is
really a circle $S^1_\beta$ parametrized by $\beta$, and similarly
the end at $w=\ell$ is a circle $S^1_\alpha$ parametrized by
$\alpha$. The rings ${\mathcal R}_q$ and ${\mathcal R}_{q'}$ are generated by
supersymmetric loop operators wrapped on $S^1_\beta$ or
$S^1_\alpha$ at a fixed time.
In generalized quiver gauge theories associated to a Riemann
surface $C$, half-BPS Wilson and 't Hooft operators are in
one-to-one correspondence \cite{DMO} with data of the following
kind: a choice of a homotopy class of simple closed loop
$\gamma\subset C$ together with a choice of a representation $R$
of $G$. Of course, this construction is usually made in
undeformed super Yang-Mills theory. However, near either end of
$S^3_{{\alpha},{\beta}}$, the vector field that we have used to
make the $\Omega$-deformation reduces to either
$\epsilon_2\partial/\partial \beta$ (at $w=0$) or
$\epsilon_1\partial/\partial\alpha$ (at $w=\ell$). So we can
rotate from the ordinary theory to the $\Omega$-deformed theory as
in section \ref{rethink}, and hence the half-BPS operators studied
in \cite{DMO} have analogs in our situation. The holonomy
functions (\ref{lovely}) that generate ${\mathcal R}_q$ and ${\mathcal R}_q'$ simply
come from the corresponding half-BPS loop operators.
It was already observed in \cite{AGT} that half-BPS loop operators
can be wrapped on the circles $S^1_\alpha$ and $S^1_\beta$ and
that in the $\Omega$-deformed theory, these operators do not
commute. What we have contributed is to relate this fact to the
noncommutativity that arises in two-dimensional $\sigma$-models
with a $B$-field, and to formulate the problem in a way that is
closer to other occurrences of the noncommutative ring ${\mathcal R}_q$ in
mathematical physics.
\subsubsection{Winding States Of The $A$-Model}\label{winda}
We now make a slight digression, aiming to spare the reader some
puzzlement by briefly answering the following question. Given the
$T$-duality between ${\mathcal B}_\beta$ and ${\mathcal B}_{{\cmmib N}}$, the space of
$({\mathcal B}_\beta,{\mathcal B}_\beta)$ strings must be equivalent to the space of
$({\mathcal B}_{{\cmmib N}},{\mathcal B}_{{\cmmib N}})$ strings. But ${\mathcal B}_{{\cmmib N}}$ is an ordinary Lagrangian
$A$-brane, so the space of $({\mathcal B}_{{\cmmib N}},{\mathcal B}_{{\cmmib N}})$ strings in the
$A$-model, as usually understood, is simply the cohomology of
${{\cmmib N}}$. How can we possibly identify the space of
$({\mathcal B}_{{\cmmib N}},{\mathcal B}_{{\cmmib N}})$ strings with a ring of holonomy functions?
The answer to this question is that it is necessary to take into
account something that is usually not relevant -- winding states
in the $A$-model. As a prototype of the problem, we consider a
sigma-model with target $W={\mathbb R}\times S^1$, with the obvious product
metric and with coordinates $t$, $\theta$ on ${\mathbb R}$ and $S^1$. First
we consider the $B$-model on $W$, in the obvious complex structure
in which $t+i\theta$ is holomorphic. Let ${\mathcal B}$ be a $B$-brane
whose support is all of $W$, with trivial Chan-Paton bundle. The
$({\mathcal B},{\mathcal B})$ strings of zero ghost number are associated to the
holomorphic functions
\begin{equation}f_n=\exp(n(t+i\theta)),~~n\in{\Bbb Z}.\end{equation}
We note that apart from $f_0$, which is the constant function 1,
all other $f_n$ have non-zero momentum around $S^1$, and
exponential growth along ${\mathbb R}$ in one direction or the other.
Now we perform $T$-duality along the $S^1$ direction. This maps
the $B$-model of $W$ to an $A$-model of $\widetilde W={\mathbb R}\times \widetilde
S^1$, where $\widetilde S^1$ is the dual circle. The brane ${\mathcal B}$ is
mapped to a Lagrangian $A$-brane ${\mathcal B}'$ that is supported on
${\mathbb R}\times p$, with $p$ a point in $\widetilde S^1$. The $({\mathcal B},{\mathcal B})$
string corresponding to the identity function $f_0$ maps to the
$({\mathcal B}',{\mathcal B}')$ string that corresponds to the zero-dimensional
cohomology of ${\mathbb R}\times p$. What about the $f_n$ with $n\not=0$?
As they carry momentum along $S^1$, they correspond to $A$-model
states that have winding around $\widetilde S^1$. Of course, these
states also have exponential growth along ${\mathbb R}$, since this
property is unaffected by $T$-duality. The reason that such
$A$-model winding states are unfamiliar is that we do not usually
study $A$-model states with exponential growth.
To see that this example really is a prototype for the original
question, note that the product of two copies of $W$ is
$X={\mathbb R}^2\times T^2$, which we can think of as the moduli space of
vacua for the familiar example of a free vector multiplet. The
product of two copies of the brane ${\mathcal B}$ considered above is the
space-filling brane ${\mathcal B}_*$ of type $(B,B,B)$ on $X$. Turning on a
Chan-Paton curvature proportional to $\omega_J$, this can be
deformed to a coisotropic $A$-brane of type $(A,B,A)$, without
changing the essentials of the above discussion. Thus, even for a
free vector multiplet, we have to go beyond the usual class of
$A$-model states to see the duality between the spaces of
$({\mathcal B}_\beta,{\mathcal B}_\beta)$ strings and of $({\mathcal B}_{{\cmmib N}},{\mathcal B}_{{\cmmib N}})$ strings.
\subsection{Partition Functions}\label{partfns}
The formulation in \cite{AGT} was actually slightly different
from what we have given here and focused on partition functions.
The main claims there were that the partition function of a
generalized quiver theory on $\R^4_{\varepsilon_1,\varepsilon_2}$ gives a chiral conformal
block in Liouville theory, and that the partition function of such
a theory on $S^4_{\varepsilon_1,\varepsilon_2}$ (at least for $\epsilon_1=\epsilon_2$) gives
the modular-invariant partition function of Liouville theory with
left- and right-movers included.
It makes sense to compare the various spaces that are involved
here, because ${\mathbb R}\times S^3$ is equivalent topologically to ${\mathbb R}^4$
with a point at the origin omitted, while $S^4$ can be viewed as
${\mathbb R}^4$ with a point at infinity added. Comparing the three spaces
in this way, the $U(1)\times U(1)$ action that we have used to
make an $\Omega$-deformation on ${\mathbb R}\times S^3$ can be extended
over ${\mathbb R}^4$ or $S^4$. Then we make the $\Omega$-deformation on
all three spaces with the same parameters
$\varepsilon_1,\varepsilon_2$, and we refer to the result as a
generalized quiver theory on ${\mathbb R}\times S^3_{\varepsilon_1,\varepsilon_2}$, $\R^4_{\varepsilon_1,\varepsilon_2}$, or
$S^4_{\varepsilon_1,\varepsilon_2}$.
In general, in any four-dimensional quantum field theory, let $B$ be a four-manifold with boundary $S^3$.
Then the path
integral on $B$ gives a physical state in the Hilbert space associated to $S^3$.
In the context of topologically twisted four-dimensional gauge theory, we have to make
a choice, as was remarked in section \ref{gl}: on ${\mathbb R}\times S^3$, topological twisting conserves
two supercharges,
$Q$ and its adjoint $\overline Q$, but only one can be conserved on a more general four-manifold
$B$. Let us make a choice and conserve $Q$.
If in addition the $U(1)\times U(1)$ action of $S^3$ extends over $B$, we can make the
$\Omega$-deformation on $B$ and in that case,
the path integral on $B$ will give a vector in the $Q$-cohomology
of the $\Omega$-deformed theory on $S^3$, in other words, a
Virasoro (or Liouville) conformal block.
The most simple choice of $B$ that has the right properties is a
four-dimensional ball $B^4$, as explained at the end of section
\ref{gl}. Writing $B^4_{\varepsilon_1,\varepsilon_2}$ for the $\Omega$-deformed version of
$B^4$, the path integral over $B^4_{\varepsilon_1,\varepsilon_2}$ will give a Virasoro
conformal block $\mathcal W$. Observing that $B^4_{\varepsilon_1,\varepsilon_2}$ is the same
topologically as $\R^4_{\varepsilon_1,\varepsilon_2}$, and that this equivalence is compatible
with the $U(1)\times U(1)$ action, what we have just said is
equivalent to the ``chiral'' version of the claim in \cite{AGT}.
For the non-chiral version of their claim, we want to glue
together two copies of $B^4_{\varepsilon_1,\varepsilon_2}$, with opposite orientation, along
their common boundary $S^3_{\varepsilon_1,\varepsilon_2}$. If we preserve the same
topological supercharge in both copies of $B^4_{\varepsilon_1,\varepsilon_2}$, we will get
Donaldson theory on $S^4_{\varepsilon_1,\varepsilon_2}$. This is not what we want. Instead,
to get the claim of \cite{AGT}, we must preserve one supercharge
$Q$ on one copy of $B^4_{\varepsilon_1,\varepsilon_2}$ and its conjugate $\overline Q$ on the other
copy. Then, the partition function on $S^4_{\varepsilon_1,\varepsilon_2}$ gives the norm
squared of the Virasoro conformal block $\mathcal W$, in the
Hilbert space $\H_{\varepsilon_1,\varepsilon_2}$.
In general, it is not clear how to make opposite topological
twists in the two hemispheres of $S^4_{\varepsilon_1,\varepsilon_2}$. For $\epsilon_1=\epsilon_2=1/R$, it has
been shown \cite{Pestun1} that the $\Omega$-deformed theory with opposite twists
on the two sides is equivalent to
physical Yang-Mills theory on a four-sphere of radius $R$,
with no $\Omega$-deformation at all. This fact was exploited in \cite{AGT}. We hope
that this fact can be adapted for our derivation and has a useful generalization to other values of the
deformation parameters. Until this is found, we may fall back on
the approach of \cite{CV} in two dimensions. In that approach,
between the two copies of $B^4_{\varepsilon_1,\varepsilon_2}$, one places a very long cylinder
$I\times S^3_{\varepsilon_1,\varepsilon_2}$. In the limit that the length of $I$ becomes much
greater than the radius of $S^3_{\varepsilon_1,\varepsilon_2}$, the path integral on
$I\timesS^3_{\varepsilon_1,\varepsilon_2}$ projects onto quantum ground states, and the path
integral on $S^4_{\varepsilon_1,\varepsilon_2}$ can be evaluated in the space $\H_{\varepsilon_1,\varepsilon_2}$ of
quantum ground states even though we make opposite topological
twists at the two ends.
\subsection{Flatness}\label{flatness}
One basic fact about conformal field theory
on a Riemann surface $C$ is that the space of conformal blocks is locally independent
of the complex structure of $C$. There is a projectively flat connection that can be used
to transport the space of conformal blocks as the complex structure of $C$ varies.
This holds for theories with finite-dimensional spaces of conformal
blocks as well as for more sophisticated theories such as Liouville theory. See \cite{FS} for more
discussion.
In our present context, this implies that although the definition
of the variety ${{\cmmib N}}$ of opers depends on the complex structure of
$C$, the space $H^0({{{\cmmib N}}},K^{1/2}_{{{\cmmib N}}})$ does not. More
precisely, the bundle of Hilbert spaces over the moduli space of
complex structures on $C$ whose fiber is $H^0({{{\cmmib N}}},K^{1/2}_{{{\cmmib N}}})$ should admit a projectively flat connection. In principle, the $\sigma$-model must generate this connection
but we do not know what sort of semiclassical formula can be given for it.
One may ask whether the underlying $(0,2)$ model in six dimensions predicts the existence of
this flat connection. The $R$-symmetry group of this theory is $SO(5)_R$. There is no non-trivial
homomorphism from $SO(6)$ to $SO(5)$, so the $(0,2)$ model does not have a twisted version
that would lead to a topological field theory in six dimensions. However, suppose that
we specialize to six-manifolds of the form ${\mathbb R}\times W$, with a product
metric; here $W$ is a five-manifold. The structure group of the tangent bundle of $W$ is
$SO(5)$, which of course does admit an isomorphism with $SO(5)_R$.
So a twisted theory can be constructed for six-manifolds of the from ${\mathbb R}\times W$. This theory
has a supercharge $Q$ obeying $Q^2=0$, and it is plausible that the cohomology of $Q$
is locally independent of the metric of $W$.
This is close to what we want. What we have been studying is an
$\Omega$-deformed version of the cohomology of the twisted theory
just described for the case $W=S^3\times C$. Of course, we did not
introduce the subject in precisely this way. Rather we started
with a four-dimensional generalized quiver theory, which can be
obtained from the $(0,2)$ theory in six dimensions by
compactifying on $C$ with a topological twist that preserves
supersymmetry. Then we compactified on $S^3$, again with a
topological twist that preserves supersymmetry. The net effect was
to compactify from six dimensions to one dimension on $W=S^3\times
C$ with a topological twist that preserves supersymmetry. There
is essentially only one way to do this -- the twist mentioned in
the last paragraph that preserves five-dimensional symmetry.
This seems promising. The only catch is that it is not clear how
to include the $\Omega$-deformation in this analysis, since one
does not have a good understanding of the $\Omega$-deformation in
six-dimensional terms. To explain the existence of a flat
connection on the bundle of Hilbert spaces $\H_{\varepsilon_1,\varepsilon_2}$ over the
Teichmuller space of $C$, we obviously need an argument that takes
the $\Omega$-deformation into account.
\subsection{Including Surface Operators}\label{surfop}
The correspondence between gauge theory and two-dimensional
conformal field theory becomes richer \cite{SurfOp,SurfOp2} if
surface operators are included. Here we consider a half-BPS
surface operator inserted at one of the two ends of
$S^3_{{\alpha},{\beta}}$. Near the support of the surface
operator, the spacetime looks like $\Sigma\times D$ where $\Sigma$
is a two-manifold and $D$ is the usual cigar geometry. The
surface operator is inserted at $\Sigma\times p$, where $p$ is the
tip of the cigar. In our applications, $\Sigma={\mathbb R}\times S^1$.
The $\Omega$
deformation is made using a vector field that generates the
rotation of $D$ around $p$ (section \ref{compom}) or one that also
acts at the same time by rotation of $S^1$ (the present section).
Far from the tip, $D$ looks like a cylinder ${\mathbb R}\times \widetilde S^1$.
A half-BPS surface operator inserted on $\Sigma\times p$ preserves
all supersymmetry that is compatible with the $\Omega$-deformation
(or with the curvature of $D$), so it can be included naturally in
our analysis. First we consider the abelian case $G=U(1)$.
We consider the simplest half-BPS surface operator. It is defined
by introducing a Lie algebra valued parameter $\alpha$ and
requiring that the curvature have a delta-function singularity
$F/2\pi \sim \alpha\delta_{\Sigma\times p}$. Equivalently, the
holonomy around the singularity is $\exp(2\pi i\alpha)$.
Now let us consider the two-dimensional sigma-model, with target
${\mathcal M}_H$, that we get by compactification on $S^1\times \widetilde S^1$.
The tip of $D$ together with the surface operator will give a
brane in this model. In the absence of the surface operator, we
observed that supersymmetry requires that the curvature along $D$
vanishes. Hence the holonomy around $\widetilde S^1$ is trivial, and
the circle-valued field that we called $A_3$ in section
\ref{compom} vanishes at the boundary. To get a more useful
description, in section \ref{otto}, we $T$-dualized this field to
another angle-valued field $\varrho$. Vanishing of $A_3$ meant
that $\varrho$ obeys Neumann boundary conditions. All other
bosonic fields in the sigma-model obey Neumann boundary conditions
for more obvious reasons. So altogether, the brane ${\mathcal B}$ that
comes from the tip of the cigar has all of ${\mathcal M}_H$ for its support.
Of course, this is familiar from section \ref{otto}.
\def{\mathcal N}{{\mathcal N}}
Including the surface operator changes this analysis in only
one way. Vanishing of curvature along $D$ now requires that the
holonomy around $\widetilde S^1$ should equal $\exp(2\pi i\alpha)$
(rather than 1) or equivalently that the boundary value of $A_3$
is equal to $\alpha$. After $T$-duality, the dual field
$\varrho$ still obeys Neumann boundary conditions, but the circle
that it parametrizes is equipped with a flat Chan-Paton line
bundle $\mathcal S$. The sole effect of the surface operator on
the brane ${\mathcal B}$ that comes from the tip of the cigar is that the
Chan-Paton bundle of that brane is tensored by $\mathcal S$. There
is no problem in describing ${\mathcal S}$ explicitly as a line bundle over
${\mathcal M}_H$. For $G=U(1)$, ${\mathcal M}_H={\mathbb R}^2\times T^2$ and ${\mathcal S}$ is simply a
flat line bundle with holonomy around one of the directions in
$T^2$. As ${\mathcal S}$ is flat, tensoring with ${\mathcal S}$ preserves all
supersymmetry.
Now let us consider the nonabelian case. We will consider the
case of a generalized quiver theory \cite{DG} based on $SU(2)$,
but we start with the case of a single $SU(2)$ gauge group. The
parameter $\alpha$ now takes values in the Lie algebra of a
maximal torus of $SU(2)$. There is really only one major change in
the above analysis. To describe it, let us work over a generic
part of the base $\cmmib B$ of the Hitchin fibration where $SU(2)$
is broken to an abelian subgroup, isomorphic to $U(1)$. Then the
monodromy around $\widetilde S^1$ is again $U(1)$-valued, computed
from the zero mode of an effective scalar field that we will again
call $A_3$. And $A_3$ is again $T$-dualized to another scalar
$\varrho$ to go to a more useful description. There is only one
major difference from the case of $G=U(1)$. Vanishing of the
curvature along $D$, instead of requiring $A_3$ and $\alpha$ to be
equal (as circle-valued objects), now requires only that they
should be equal up to a Weyl transformation. Thus (for
$G=SU(2)$), there are now two allowed boundary values of $A_3$,
namely $A_3=\pm\alpha$.
\def{\mathcal U}{{\mathcal U}}
After $T$-duality, each of the allowed boundary values of $A_3$
maps to a flat line bundle over the circle parametrized by
$\varrho$. As $A_3$ has two allowed boundary values, this gives a
rank two flat bundle over the circle. As the other fields vary,
this will give a rank two bundle ${\mathcal U}\to{\mathcal M}_H$ (at least in the
region where our approximations apply). The effect of the surface
operator is to tensor the Chan-Paton bundle of the brane ${\mathcal B}$ that
comes from the tip of the cigar with ${\mathcal U}$.
Our description so far has been based on an approximation valid
far from the discriminant locus in $\cmmib B$. In this
approximation, ${\mathcal U}$ is a flat vector bundle. It is not possible
for this to be the whole story, since the total space of the
Seiberg-Witten or Hitchin fibration for any gauge theory with a
semi-simple gauge group has a finite fundamental group. The most
optimistic hypothesis is that the full description of the surface
operator involves a rank two bundle ${\mathcal U}\to{\mathcal M}_H$ that is flat far
from the discriminant locus and in general preserves all
supersymmetry. We will assume this to be the case, though it is
also imaginable that something more complicated happens near the
discriminant locus in $\cmmib B$.
Before making a proposal for a more concrete description of ${\mathcal U}$,
we will consider $SU(2)$ generalized quiver theories that arise
from six dimensions by compactification on a Riemann surface $C$.
The gauge group is a product of $SU(2)$ factors. According to
\cite{SurfOp}, in such a theory, there is a natural family of
surface operators parametrized by the choice of a point $q\in C$.
In suitable local descriptions, an operator in this family can be
described precisely as above with the relevant part of the gauge
group being simply $SU(2)$. Hence the above analysis applies, and
the insertion at the tip of the cigar of a surface operator
associated to a point $q\in C$ will generate a rank two bundle
${\mathcal U}_q\to {\mathcal M}_H$.
As ${\mathcal U}_q$ varies in $C$, the ${\mathcal U}_q$ will fit together as fibers of
a rank two holomorphic vector bundle ${\mathcal U}\to{\mathcal M}_H\times C$. There
is an obvious candidate for what this bundle may be: the universal
bundle, or more exactly, the bundle part of the universal Higgs
bundle. In other words, every point $r\in {\mathcal M}_H$ parametrizes a
solution of Hitchin's equations -- a holomorphic bundle $E\to C$
together with a Higgs field $\varphi\in
H^0(C,{\mathrm{ad}}(E)\otimes K_C)$ and a hermitian metric such
that Hitchin's equations are satisfied. (Here we emphasized a
holomorphic point of view, but of course there are many equivalent
descriptions of this data.) The universal Higgs bundle is a triple
consisting of a holomorphic bundle ${\mathcal E}\to{\mathcal M}_H\times C$,
a Higgs field $\widehat\varphi:{\mathcal E}\to {\mathcal E}\otimes K_C$,
and a hermitian metric on ${\mathcal E}$, such that the
restriction of this data to $r\times C$, for any $r\in {\mathcal M}_H$, is
the solution of Hitchin's equations corresponding to $r$. (For an
elementary description of universal bundles, including subtleties
involving the center of the gauge group, see section 7.1 of
\cite{KW}.) By restricting $\mathcal E$ to ${\mathcal M}_H\times q$, we get
a very plausible candidate for ${\mathcal U}_q$.
In the absence of the $\Omega$ deformation, the condition that
tensoring with ${\mathcal U}_q$ preserves all supersymmetry means that
${\mathcal U}_q$ must be holomorphic in every complex structure on ${\mathcal M}_H$.
This follows from standard properties of the universal Higgs
bundle. Upon making the $\Omega$ deformation, further properties
are needed, and they are not well-understood, since the theory of
coisotropic $A$-branes of rank greater than 1 has not been
developed. Possibly the necessary properties follow from the
existence and properties of the universal Higgs field
$\widehat\varphi$.
\vskip 1 cm
\noindent{\it Acknowledgments}
Part of this work has been done while NN visited the IAS and while
EW visited the IHES and the SCGP. We thank these institutes for their hospitality.
Research of NN was partly supported by {\it l'Agence Nationale de la
Recherche} under grants
ANR-06-BLAN-3$\_$137168 and ANR-05-BLAN-0029-01, and by the Russian
Foundation for Basic Research through the grants
RFBR N 09-02-00393 and NSh-3036.2008.2.
Research of EW was partly supported by
NSF Grant Phy-0503584. We thank R. Donagi, E. Frenkel, D.
Gaiotto, S. Gukov, G. Moore, S. Shatashvili, J. Teschner, and Y.
Tachikawa for comments and discussions.
|
\section{\label{}}
\section{Introduction}
Recent observational progresses revealed that
Gamma-ray Bursts (GRBs) are extremely energetic explosion
originated by core collapse of the massive stars or
merger of binary of neutron stars or black holes at
cosmological distance. The gamma-ray prompt emission below
several MeV band has been usually observed but the detection
of high-energy emission above 100 MeV was reported only a
few times by the Energetic Gamma-Ray Experiment Telescope (EGRET)\cite{Dingus}
and recently by Astro-rivelatore Gamma a Immagini LEggero (AGILE)\cite{Giuliani}.
The observed properties of these high-energy photons from some GRBs
showed distinct spectral and temporal behavior compared with low
energy photons below several MeV, suggesting different gamma-ray
emission processes between low and high-energy photons\cite{Hurley}\cite{Gonzaretz}.
In addition, comparing observed properties of high-energy photons
between short and long duration GRBs will help to study the sub
classes of GRBs. However, very low sample number of
GRBs with high-energy photons before Fermi era makes it difficult to
reveal the detailed properties of high-energy emission from GRBs.
The Fermi Gamma-ray Space Telescope has successfully detected 14 GRBs
with high-energy photon ($>$100 MeV) so far thanks to its unprecedented effective
area above 100 MeV of the Large Area Telescope (LAT)\cite{Atwood}.
Figure\ref{grb_skymap} shows the
sky distribution of all GRBs detected by the LAT and Gamma-ray Burst Monitor (GBM)
as of 22th January, 2010.
Fermi firmly confirmed the distinct high-energy spectral component with
respect to traditional Band function in low energy band observed by the
GBM from some bright GRBs so far (GRB 090510 and GRB 090902B). Furthermore, high-energy photon
detected by short-hard GRB 090510 provides a stringent photon dispersion limit,
which strongly disfavors the quantum gravity model of space-time causes a linear
variation of the speed of light with photon energy\cite{grb090510}.
In addition to those interesting properties for individual LAT GRBs, the large
number of samples enable us to study the systematic properties of
high-energy emission from GRBs. Actually, the high-energy photons of the LAT
are often delayed and extended compared with the low-energy emission from GRBs,
suggesting a different acceleration process and/or emission process in high
energy emission. And also Fermi has detected high-energy photons from both
short and long duration GRBs. This would help to classify the sub classes of
GRBs with high-energy emission properties.
In this paper, we present the detail analysis result of three GRBs with
weak LAT detection, GRB 080825C\cite{grb080825c}, GRB 081024B\cite{grb081024b},
and GRB 090217\cite{grb090217} and discuss
the global properties of the high-energy emission of GRBs by comparing with
other LAT bright GRBs.
\begin{figure}
\rotatebox{-90}{\includegraphics[width=40mm]{figure/figure1.eps}}
\caption{GRB skymap detected by Fermi between 14th July, 2008 to 22th January, 2010. GRBs detected by
the GBM in field of view of the LAT and out of
field of view of the LAT are shown by red and black asterisks, respectively.
GRBs detected $>$100 MeV photons by the LAT are
shown by large blue asterisks}
\label{grb_skymap}
\end{figure}
\section{Analysis Techniques}
We have performed a detailed temporal and spectral analysis for three GRBs with weak
LAT emission through the careful procedures, which were developed by our collaborations
for event selection, selection of the region of interest (ROI), and background
estimation. In this section, we briefly summarize these
important procedures for our analysis. See Abdo et al. 2009a for more detail.
\subsection{Event Selection}
The standard selection of the LAT photon event has been developed by Atwood et al. 2009.
There are three event classes, named as ``diffuse'', ``source'', and ``transient''
class for specific scientific analysis. In the case of GRB observations, the smaller region of the
sky and shorter time scale compared with diffuse sources allows the event section to be
relaxed, and the ``transient'' event class is usually used to detection, localization, and
analysis for the prompt emission. On the other hand, diffuse class, which is used to analysis
a faint source with longer time interval and covering larger region of the sky
is also used for the afterglow search of GRBs. In addition to these standard event classes,
we developed more optimized event selection than the ``transient'' class for the spectral
analysis based on Monte-Carlo simulations. From this study, we found that more relaxed can be
used because the background contamination is less issue due to short time windows of GRB analysis.
This optimized event class, so-called ``S3'' class is used for the spectral analysis of
GRB 080825C. For other GRBs, GRB 081024B and 090217, the standard ``transient'' class is applied
for the spectral analysis.
\subsection{Energy-dependent ROI}
The LAT point-spread function (PSF) strongly depends on the incident energy and the conversion
point on the tracker, and thus the detector response of the LAT is separated into ``FRONT'' and
``BACK'' events. We also consider the energy-dependent region of interested (ROI) for these ``FRONT''
and ``BACK'' event individually based on the 95 \% containment radius (PSF95) and the 95 \% LAT
localization error (Err95) as follows:
\begin{equation}
ROI(E) = \sqrt{PSF95(E)^2+Err95^2}
\end{equation}
The maximum value of this energy-dependent ROI is set to 10 and 12 degree for ``FRONT'' and ``BACK'' event, respectively not to contain too large background region. We apply this energy-dependent ROI
for all three GRBs.
\subsection{Background Estimation}
Because of very small number of photons detected for three GRBs presented here, the
background estimation should be performed very cautiously. The background rate
strongly depends on many parameters such as the incident angle of the burst or
the position in the instrument.
Therefore, it is not straightforward to estimate the accurate LAT background using
off-source region around the trigger time.
There are two components of LAT background events;
cosmic-ray (CR) background and diffuse gamma-ray background (from extra-galactic and Galactic).
Since these components have different properties, we have developed two different methods to estimate
the expected background rate. The amount of gamma-ray background only depends on the exposure
of a certain direction in the celestial sphere. Therefore, this component can be estimated
by a simple scaling of the number of gamma-rays detected in longer observations. Here, we
used six months of LAT data for this scaling. The amount of CR background also depends on
the exposure of specific direction of the sky, however, this component also changes with the
geomagnetic coordinates at the location of the spacecraft, and this dependence cannot be
estimated by the same way as the gamma-ray background. Thus, we utilized Monte Carlo simulation
of the GRB observation to estimate the CR background.
\section{Analysis Results of Three LAT GRBs}
\subsection{Detection and Localization}
GRB 080825C is the first long GRB detected by the LAT\cite{Bouvier}.
The GBM triggered on this burst at 14:13:48 UT\cite{van der Horst}.
We estimated the burst location using ``transient'' class described above.
In this analysis, we fit the LAT data assuming the power-law shape of the
point source. To estimate the error of the position of this point source,
we calculate the test statistics (TS), assuming point source for each grid
of the map. As the result of this analysis, we found the best-fit position
of this GRB at (RA,Dec)=(233.9, -4.5) with
0.8 degree of 68 \% error radius. The detection significance is estimated by
several independent method such as unbinned likelihood analysis or some Bayesian
approaches, and we obtained about 6 sigma by these methods. Detailed description
of these methods can be found in Abdo et al. 2009a.
The short GRB 081024B triggered the GBM at 21:22:41 UT\cite{Connaughton} and the LAT ground
analysis found high-energy photons from this burst up to 3 GeV\cite{Omodei}. Thus, this
is the first short GRBs which is detected GeV photon.
Through the same procedure as GRB 080825C using ``transient'' class events,
the LAT location of this GRB was estimated as (RA,Dec)=(322.86, 21.16) with 0.22
degree of 68 \% error, and
the detection significance of this burst is 6.7 sigma.
Another long GRB, 090217 triggered the GBM at 04:56:42 UT\cite{von Kienlin}. This burst firstly
detected by the blind search of the LAT data by the on-ground
Automated Science Processing (ASP), and this is confirmed by the follow-up analysis
around GBM position\cite{Ohno}. The ``transient'' class events are used to obtain the location and
significance of the LAT data same as GRB 080825C, and the best-fit position is found to
be (RA, Dec)=(204.74, -8.43) with 0.37 degree error radius. The detection significance is
8.4 sigma and 9.2 sigma with the semi-Bayesian method and maximum TS value obtained by
unbinned likelihood fit, respectively.
\subsection{Temporal Properties}
Figure \ref{latgbmlc} shows the energy-resolved light curve of three LAT GRBs.
The top two panels in each figure show the background-subtracted light curve
of NaI and BGO detectors of GBM. The LAT events around the GBM position
are plotted on the other bottom panels in each figure.
From these light curves, we can see various temporal properties of high-energy photons
from bursts to bursts. First, we can clearly see that the onset of LAT high-energy photons
are delayed with respect to the GBM low energy photons for GRB 081024B. In the case of
GRB 080825C, this delay can be seen but it is not statistically significant.
Furthermore, the LAT high-energy photons last longer than GBM low energy emission;
$\sim$35 s for GRB 080825C, and $\sim$3 s for GRB 081024B. The highest energy photon, 572 MeV
for GRB 080825C and 3.1 GeV for GRB 081024B are detected when the low energy emission becomes
weak. Whereas the LAT high-energy photons shows various properties different from GBM
low energy photons for GRB 080825C and GRB 081024B, few noticeable temporal features can be
observed from GRB 090217: there is no delay and extended emission of high-energy photons.
\subsection{Spectral Properties}
The time-resolved spectral analysis combining the LAT and GBM
data was performed for three weak LAT GRBs.
Figure \ref{latgbmspec} shows the resulting spectral models in the
$\nu$F$_\nu$ representation. Each figure shows the
best-fit model in each time interval shown in figure \ref{latgbmlc}.
From this analysis, we found no significant evidence of an additional
high-energy spectral component such as extra power-law component or
high-energy cut-off structure, and the single
Band function or Comptonized model give a good fit for both time-integrated
and time-resolved spectrum for all three weak LAT GRBs. We only found a weak
evidence of high-energy cut-off from GRB 080825C, E$_{\rm cut}=1.77^{+1.59}_{-0.56}$ MeV
at the first time bin with the significance level of 4.3 sigma.
Interestingly, significant spectral hardening at the high-energy band in the last
time bin can be seen for GRB 080825C and GRB 081024B, where the GBM low energy
emission is almost back to background level. The spectrum of this late-time high
energy emission can be represented by a single power-law model with the photon index of
$-1.95\pm0.05$ and $-1.59^{+0.2}_{-0.07}$ for GRB 080825C and GRB 081024B, respectively.
This late-time high-energy component is common to other bright LAT GRBs and the possible
origin of this emission is synchrotron self-Compton emission during
afterglow phase or cascades induced by ultrarelativistic hadrons accelerated by the
relativistic jet.
GRB 090217, however, shows no significant evidence of additional spectral component and
spectral evolution in any time interval unlike GRB 080825C and GRB 081024B.
\section{Comparison with other bright LAT GRBs}
Table \ref{grbcomp} shows the comparison of the various properties between weak and bright LAT
GRBs. Above temporal and spectral analysis for three weak LAT GRBs revealed that
the delayed onset and long-lived behavior of high-energy photons from
GRB 080825C and GRB 081024B. Such behavior is also reported from other bright
LAT GRBs such as GRB 080916C, 090510, and 090902B, and we might think that
the delayed onset and long-lived high-energy photon is the common feature of
GRBs. However, we can not find any temporal and spectral feature from GRB 090217.
Such ``featureless'' feature of GRB 090217 may suggest the unique mechanism of the
broad band gamma-ray emission or indicate that the different high-energy emission
class exist. When we compare the high-energy emission properties between short and
long duration GRBs, there is no clear difference between these two classes in both
weak and bright LAT GRBs.
More GRB observations are needed to investigate common and uncommon properties and
possible classification in high-energy emission of GRBs, and such study would be
important to give new insight on the acceleration mechanisms for the high-energy emission
of GRBs.
\clearpage
\begin{figure}[!tbp]
\centering
\includegraphics[width=80mm]{figure/figure2a.eps}
\includegraphics[width=80mm]{figure/figure2b.eps}
\includegraphics[width=80mm]{figure/figure2c.eps}
\caption{Energy-resolved light curves of three weak LAT GRBs. Top two panels are background subtracted
GBM light curve and the other bottom panels are the LAT light curves. The LAT events are the ``S3'' events above 80 MeV for GRB 080825C,
all ``transient'' events and ``transient'' events above 100 MeV for GRB 081024B, and
``transient'' events above 50 MeV and above 100 MeV for GRB 090217, respectively. Black dots in each figure show the energy of each LAT event with 1 $\sigma$ error. The vertical dashed-line in each figure shows the time-interval used in the time-resolved spectral analysis.}
\label{latgbmlc}
\end{figure}
\begin{figure}[!h]
\centering
\includegraphics[width=75mm]{figure/figure3a.eps}
\includegraphics[width=82mm]{figure/figure3b.eps}
\includegraphics[width=85mm]{figure/figure3c.eps}
\caption{The best-fit models obtained by the time-resolved spectral analysis for three weak LAT GRBs.
Solid lines in each figures show the best-fit model. The dashed-lines around these best-fit models represent the contour level at the 68 \% confidence interval. Each solid lines corresponds to time-resolved spectra which are extracted in the time interval shown in figure \ref{latgbmlc}.}
\label{latgbmspec}
\end{figure}
\begin{table*}[t]
\begin{center}
\caption{Comparison of properties of high-energy (HE) photons of weak LAT GRBs with other bright LAT GRBs.}
\begin{tabular}{|c|c|c|c|c|c|c|c|c|}
\hline
GRB & short/long & num of events & num of events& delayed & long-lived & extra comp. & highest energy & redshift \\
& class & $>$ 100 MeV & $>$ 1 GeV &HE onset & HE emission & & & \\
\hline
\hline
\multicolumn{9}{|c|}{weak LAT GRBs}\\
\hline
080825C\cite{grb080825c} & long & $\sim$ 10 & 0 & $\circ$ & $\circ$ & x & $\sim$ 572 MeV & x\\
081024B\cite{grb081024b} & short & $\sim$ 10 & 2 & $\circ$ & $\circ$ & x & $\sim$ 3.1 GeV & x \\
090217\cite{grb090217} & long & $\sim$ 10 & 0 & x & x & x & $\sim$ 866 MeV & x \\
\hline
\hline
\multicolumn{9}{|c|}{bright LAT GRBs}\\
\hline
080916C\cite{grb080916c} & long & $>$ 100 & $>$ 10 & $\circ$ & $\circ$ & $\circ$(weak) & 13.2 GeV & 4.35 \\
090510\cite{grb090510}\cite{grb090510physics} & short & $>$ 150 & $>$ 20 & $\circ$ & $\circ$ & $\circ$ & 31 GeV & 0.903 \\
090902B\cite{grb090902b} & long & $>$ 200 & $>$ 30 & $\circ$ & $\circ$ & $\circ$ & 33 GeV & 1.822 \\
\hline
\end{tabular}
\label{grbcomp}
\end{center}
\end{table*}
\bigskip
|
\section{Introduction}
It is generally accepted that the first-order isotropic-to-nematic (IN)
transition in liquid crystals confined between two parallel plates becomes
continuous when the distance $H$ between the plates becomes small
\cite{citeulike:4006298, citeulike:4067023, citeulike:3687059,
citeulike:3687084, citeulike:4066932, citeulike:4057691}. Indeed, many
simulations are consistent with this picture \cite{cleaver.allen:1993,
lagomarsino.dogterom.ea:2003, dijkstra.roij.ea:2001, citeulike:4922416,
citeulike:3683414} and show that the first-order IN transition terminates at a
critical film thickness $H_{\rm x}$. Some of these studies have also
provided evidence of a continuous transition taking place when $H<H_{\rm x}$. Note
that as $H \to 0$ the system becomes effectively two-dimensional (2D). More
recently, a (mathematically rigorous) proof appeared, showing that first-order
IN transitions in 2D are also possible \cite{physrevlett.89.285702,
enter.romano.ea:2006}. Inspired by this proof, computer simulations of liquid
crystals in 2D were performed, which indeed uncovered strong first-order IN
transitions too \cite{vink:2006*b, vink.wensink:2007}. Hence, the IN transition
in confinement can be continuous, as well as first-order. Finally, there is the
scenario of no transition occurring at all in thin films
\cite{citeulike:5158181, citeulike:5158109}, not even a continuous transition of
the Kosterlitz-Thouless (KT) type \cite{kosterlitz.thouless:1972}. Regarding
experiments on confined liquid crystals, it has proved difficult to resolve
continuous IN transitions in thin films \cite{citeulike:4067023,
citeulike:4006159}. Pronounced coexistence between isotropic and nematic
domains is typically observed \cite{citeulike:4006159, citeulike:2811025,
citeulike:3991276}, which suggests that a transition does take place and that
it is first-order.
The qualitatively different manifestations of the IN transition in confinement
(continuous, first-order, absence) rule out any universality class for this
transition. What remains of the IN transition in thin films is determined by
microscopic detail. The only regime where some \ahum{agreement} may be obtained
is in the bulk 3D limit $H \to \infty$. Here, one usually observes a first-order
IN transition, with long-range order in the nematic phase. The transition thus
breaks the rotational symmetry of the isotropic phase. At the mean-field level,
this implies that the transition must be first-order \cite{citeulike:6170476}.
We emphasize that fluctuations can change this result: even in 3D bulk, a
genuine {\it continuous} IN transition is also possible \cite{citeulike:6170476,
citeulike:6581608}. However, most bulk experiments yield a first-order IN
transition, and so the mean-field approximation appears to be valid in
this regime. As the film thickness $H$ decreases, fluctuations become
increasingly important, and we expect three scenarios to unfold. In the first
and most commonly accepted scenario, the IN transition becomes continuous when
the film thickness drops below a critical thickness $H_{\rm x}$. In addition,
confinement is expected to destroy long-range order in the nematic phase, due to
the Mermin-Wagner theorem \cite{physrevlett.17.1133}. Instead, quasi-long-range
order may result, where the orientational correlations decay as a power law with
distance. In the second (lesser known) scenario, the IN transition remains
first-order irrespective of the film thickness, i.e.~all the way down to $H \to
0$. In the third scenario, evidence for which was recently provided
\cite{citeulike:5158181, citeulike:5158109}, the IN transition vanishes
completely in the thin-film limit.
Given the three scenarios for the IN transition in confinement, all of which are
qualitatively different, it is of fundamental interest to establish which
\ahum{microscopic detail} is responsible for the scenario that ultimately
occurs. The aim of this paper is to identify one possible mechanism, using
computer simulations of a generalized Lebwohl-Lasher (LL) model. As it turns
out, the generalized LL model is capable to reproduce all three scenarios, by
tuning just a single parameter in the Hamiltonian. The effect of this parameter
is to make the pair interaction \ahum{sharp and narrow}, meaning that particles
interact when aligned but are otherwise rather indifferent to each other.
Depending on this parameter, the crossover with decreasing film
thickness from first-order to continuous behavior can be eliminated completely,
and the IN transition remains first-order irrespective of $H$.
The outline of this paper is as follows. We first introduce the generalized LL
model and describe the simulation method. Next, we present new simulation data
showing one example where the IN transition becomes continuous below a critical
film thickness $H_{\rm x}$, and a second example where the transition remains
first-order irrespective of the film thickness. We do not consider the scenario
where the transition vanishes below $H_{\rm x}$ as this has recently been done
elsewhere \cite{citeulike:5158181, citeulike:5158109}. A stringent test of the
Kelvin equation, describing the shift of the transition temperature as a
function of film thickness is also included. We end with a discussion
and summary in \sect{the_end}.
\section{model and simulation method}
\label{model}
We consider a lattice model similar in spirit to the LL~model
\cite{physreva.6.426}. To each site~$i$ of a 3D lattice, a 3D unit vector
$\vec{d}_i$ (spin) is attached, which interacts with its nearest neighbors via
\begin{equation}\label{eqll}
E = - \epsilon \sum_{\langle i,j \rangle} | \vec{d}_i \cdot \vec{d}_j |^p,
\end{equation}
with exponent $p$ and coupling constant $\epsilon$. In this work we absorb a
factor of $1/k_B T$ in the coupling constant, with $k_B$ the Boltzmann constant
and $T$ the temperature, and so $\epsilon$ plays the role of inverse
temperature. The lattice is a $L \times L \times H$ rectangular box, with
periodic boundary conditions in the lateral $L$ directions but not in the $H$
direction. The parameter $H$ thus plays the role of the film thickness; the
minimum thickness that can be studied in this way equals $H=1$, corresponding to
a single lattice layer. This setup is identical to the slab geometry used in
earlier simulations of the confined LL~model \cite{cleaver.allen:1993}. Note
that spins at the walls have a lower number of nearest neighbors, but that the
walls are otherwise neutral, i.e.~we do not impose any anchoring conditions.
In the original LL~model the exponent of \eq{eqll} equals $p_{\rm LL}=2$. In the
bulk limit $H \to \infty$, a (weak) first-order IN transition is observed
\cite{citeulike:5091146, citeulike:3740162, physreva.6.426, citeulike:4197190,
physrevlett.69.2803} at $\epsilon_\infty \approx 1.34$ \cite{note1}. In the
thin-film limit $H=1$ recent simulations indicate the absence of any transition
when $p=2$ \cite{citeulike:5158181, citeulike:5158109}. In this work we consider
$p>2$. This modification is expected to enhance first-order phase transitions
\cite{citeulike:5091592}, which may then even survive the limit $H \to 1$
\cite{vink:2006*b, vink.wensink:2007}. In line with previous work
\cite{citeulike:5202797, physrevlett.69.2803, citeulike:3740162}, we analyze
\eq{eqll} in terms of the histogram
\begin{equation}\label{eq:histo}
P(E,S) \equiv P(E,S | H,L,\epsilon),
\end{equation}
defined as the probability to observe a system with energy $E$ and nematic order
parameter $S$ in a sample of thickness $H$, lateral extension $L$ and at inverse
temperature $\epsilon$. The distribution is obtained by computer simulations
using Wang-Landau \cite{wang.landau:2001, citeulike:278331} and transition
matrix \cite{citeulike:202909} sampling; additional details pertaining to the
present model are provided in \olcite{citeulike:5202797}. The nematic order
parameter $S$ is defined in the usual way as the maximum eigenvalue of the
orientational tensor
\begin{equation}
Q_{\alpha \beta} = \frac{1}{2N} \sum_{i=1}^N
\left( 3 d_{i\alpha} d_{\beta} - \delta_{\alpha\beta} \right),
\end{equation}
with $d_{i\alpha}$ the $\alpha$ component ($\alpha=x,y,z$) of the orientation
$\vec{d}_i$ of the spin at site $i$, the sum over all $N=HL^2$ lattice sites
and $\delta_{\alpha\beta}$ the Kronecker delta. In a perfectly aligned sample
it holds that $S=1$, whereas an isotropic sample yields $S \to 0$ in the
thermodynamic limit (hence, $S$ defined in this way is an intensive quantity).
A final ingredient of this work is the use of finite-size scaling (FSS); needed
because we seek thermodynamic limit properties. The thermodynamic limit of a
film of thickness $H$ is defined by extending the lateral extension $L \to \infty$.
In the bulk thermodynamic limit, both $H$ and $L$ are taken to infinity.
\section{Results}
Depending on the exponent $p$ in \eq{eqll}, we expect the first-order IN
transition either to terminate at a critical film thickness $H_{\rm x}$ or to remain
first-order irrespective of $H$. The case $p=2$, i.e.~the original LL model, is
an example of the former scenario. In the bulk limit one obtains a first-order
transition \cite{citeulike:5091146, citeulike:3740162, physreva.6.426,
citeulike:4197190, physrevlett.69.2803}, which terminates when the film
thickness equals $H_{\rm x} \sim 8 - 16$ lattice layers \cite{cleaver.allen:1993}. In
the 2D limit $H=1$ no phase transition is observed for $p=2$
\cite{citeulike:5158181, citeulike:5158109}. We emphasize that the latter
finding is not without some controversy, as previous other numerical studies of
this system concluded that a phase transition does take place, namely a
continuous transition of the KT type (see discussion in
\olcite{citeulike:5158181}).
\subsection{crossover scenario}
\begin{figure}
\begin{center}
\includegraphics[width=1.0\columnwidth]{op_dis}
\caption{\label{fig:op_dis} Logarithm of $P$ using $p=20$ in \eq{eqll} with
$H=1$ and $L=25$. The value of $\epsilon$ has been chosen to give peaks of
equal height. The free energy barrier, labeled $\Delta F$, is given as the
difference between the peak maxima straddling the minimum. The distance
labeled $\Delta \rho$ corresponds to the latent heat density. The distribution
is plotted as a function of the negative energy density, such that the left peak
corresponds to the isotropic phase and the right peak to the nematic phase.}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=1.0\columnwidth]{xover}
\caption{\label{fig:xover} Evidence of the crossover scenario, whereby the IN
transition ceases to be first-order below a critical film thickness; the results
in this plot refer to $p=8$ in \eq{eqll}. The free-energy barrier $\Delta F$
versus the lateral film extension $L$ is plotted for several values of the film
thickness $H$ in (a). For large $H$ the barriers increase linearly with $L$,
consistent with a first-order transition; for smaller $H$ the barrier vanishes
with increasing $L$. The maximum value of the specific heat
versus $L$ is plotted in (b), again for several $H$.}
\end{center}
\end{figure}
We now consider \eq{eqll} using a larger exponent, $p=8$, to demonstrate that
also a continuous IN transition is possible in thin films. To determine the
order of the transition we use two FSS methods: the first was initially given
by Lee and Kosterlitz \cite{physrevlett.65.137} and is based on the energy
distribution $P(E)$, defined as the probability to observe a system with energy
$E$
\[
P(E) \equiv \int \int \delta(E-E') P(E',S') dE' dS',
\]
with $P(E,S)$ the joint distribution of \eq{eq:histo}.
At a first-order transition $P(E)$ becomes bimodal, see \fig{fig:op_dis} for
an example, where the {\it logarithm} of the distribution is shown. For
finite $L$ the bimodal structure persists over a range of $\epsilon$ values.
As $L$ increases the range becomes smaller and in the thermodynamic limit
$L \to \infty$ there is only one $\epsilon$ where $P(E)$ is bimodal, then
featuring two $\delta$-peaks. Hence, for finite $L$ there is some freedom
in choosing $\epsilon$ and in \fig{fig:op_dis} we have tuned $\epsilon$
such that the peaks are of equal height.
At a first-order transition the peak height, $\Delta F$ in $\ln P(E)$, see the
vertical arrow in \fig{fig:op_dis}, corresponds to the free energy cost of
interface formation \cite{binder:1982}. We therefore expect $\Delta F \propto
L^{d-1}$, with $L$ the lateral extension of the film and $d=2$ (recall that
films are effectively two-dimensional). To determine the order of the
transition, Lee and Kosterlitz \cite{physrevlett.65.137} proposed to measure
$\Delta F$ versus $L$, which should yield a {\it linear} increase for a film.
Results for several values of $H$ are shown in \fig{fig:xover}(a). The data
clearly indicate that the crossover scenario is taking place: for $H=6$ $\Delta
F$ increases linearly with $L$, consistent with a first-order transition. In
contrast, for $H=3$ $\Delta F$ vanishes for large $L$, implying the absence of a
first-order transition.
To obtain the crossover thickness $H_{\rm x}$ more accurately we use a second FSS
method, based on the specific heat
\begin{equation}\label{eqcv}
C = (\avg{E^2} - \avg{E}^2)/N,
\end{equation}
with $N=HL^2$ the number of lattice sites (volume). For given $L$ and $H$ a
graph of $C$ versus $\epsilon$ reveals a maximum; the value of the maximum
defines $C_{L, \rm max}(H)$. At a first-order transition the maximum scales with
the volume of the system, that is $C_{L, \rm max}(H) \propto N$
\cite{citeulike:3610966}. In a film of fixed thickness $H$ this implies $C_{L,
\rm max}(H) \propto L^{\tilde\alpha}$ with ${\tilde\alpha}_{\rm 1st}=2$. The result is shown
in \fig{fig:xover}(b) for several values of the film thickness. For $H=6$ a fit
yields $\tilde\alpha=2.00$, confirming that the transition is first-order. For
$H=4$ we obtain $\tilde\alpha=1.74$, indicating that a first-order transition is
absent. Hence, we conclude that the crossover thickness $H_{\rm x}=5$. Precisely at
$H_{\rm x}$ a fit yields $\tilde\alpha=1.94$, which is still very close to the
first-order value. Presumably for $H=5$ the IN transition is weakly first-order.
\begin{figure}
\begin{center}
\includegraphics[width=8cm]{op}
\caption{\label{op} Variation of the nematic order parameter $S$ versus inverse
temperature $\epsilon$ using $p=8$ in \eq{eqll} for several values of the film
thickness $H$. In (a) we show the bulk result $H \to \infty$, whereas (b) and
(c) were obtained in films of finite thickness $H$. Note that for (a) and (b)
the IN transition is first-order while it has become continuous in (c).}
\end{center}
\end{figure}
For $H>H_{\rm x}$, i.e.~where the transition is distinctly first-order, there is
two-phase coexistence at the transition inverse temperature. It seems natural to
characterize the phases with the nematic order parameter $S$. This approach is
somewhat dangerous as confinement could destroy long-range nematic order in the
thermodynamic limit: $\lim_{L \to \infty} S=0$ irrespective of $\epsilon$. For
$H=1$ this follows rigorously from the Mermin-Wagner theorem
\cite{physrevlett.17.1133}. The practical problem, affecting both simulations
and experiments \cite{bramwell.holdsworth:1994}, is that the decay of $S$ with
$L$ may be very slow. In fact, finite samples at low temperature typically
reveal substantial order, even when the Mermin-Wagner theorem applies
\cite{bramwell.holdsworth:1994}. The present simulations are no exception. Shown
in \fig{op}(a) is $S$ versus $\epsilon$ in the bulk limit $H \to \infty$ for
several system sizes $L$ (the bulk simulations were performed on a 3D cube of
edge $L$ with periodic boundaries in all directions). A first-order IN
transition taking place at $\epsilon \approx 1.52$ \cite{citeulike:5202797},
where $S$ jumps to a finite value, is clearly seen. More importantly, for
$\epsilon$ above the transition, $S$ becomes independent of system size, at
least on the scale of the graph; the latter is consistent with the formation of
long-range nematic order, as expected in 3D. In \fig{op}(b) we show the
corresponding result for a film of thickness $H=10$, which is still above the
crossover thickness, and so the transition remains first-order. The behavior is
similar to the bulk case, in the sense that $S$ \ahum{jumps} at the transition,
and for large $\epsilon$ it appears to saturate at a finite value independent of
the lateral film extension $L$. Hence, \fig{op}(b) provides no evidence of $S$
decaying to zero in the thermodynamic limit $L \to \infty$, but rather that the
film supports long-range nematic order. If $S$ eventually does decay to zero, it
is clear that huge system sizes, beyond the reach of any foreseeable simulation,
are required to observe it.
\begin{figure}
\begin{center}
\includegraphics[width=1.0\columnwidth]{phase_p8}
\caption{\label{phase} Capillary phase diagram of \eq{eqll} using $p=8$. Shown
is the variation of the coexisting phase densities $\rho_{\rm iso}(H)$ and
$\rho_{\rm nem}(H)$ with the inverse film thickness $1/H$. The critical inverse
thickness is at $1/H_{\rm x} \sim 0.2$, above which the transition is no longer
first-order and hence the two branches terminate. In the region between both
branches coexistence between isotropic and nematic phases is observed.}
\end{center}
\end{figure}
To avoid these subtleties, we characterize the coexisting isotropic and nematic
phases in the film with their energy densities $\rho_{\rm iso}(H)$ and
$\rho_{\rm nem}(H)$ respectively. These are simply the peak positions
in the energy distribution of \fig{fig:op_dis}. Recall that the latent heat of
the transition equals ${\cal L}_L(H) = \rho_{\rm nem}(H) - \rho_{\rm iso}(H)$,
where the subscript is a reminder of finite-size effects in the
lateral film extension. The latent heat is related to the specific heat maximum
\cite{citeulike:3610966}
\begin{equation}\label{eq:cvlh}
{\cal L}_L(H) = \sqrt{4 C_{L, \rm max}(H) / N}
\end{equation}
and the extrapolation to $L \to \infty$ is performed assuming that ${\cal
L}_\infty(H) - {\cal L}_L(H) \propto 1/N$. The average energy density
\[
\rho_L(H) \equiv
\frac{\rho_{\rm iso}(H) + \rho_{\rm nem}(H)}{2} =
\frac{1}{N} \int E P(E) \, d E
\]
obtained at the specific heat maximum is extrapolated analogously:
$\rho_\infty(H) - \rho_L(H) \propto 1/N$. Once ${\cal L}_\infty(H)$ and
$\rho_\infty(H)$ have been determined, the coexisting energy densities follow.
The latter may then be plotted in a capillary phase diagram, see \fig{phase},
where the coexistence densities versus inverse film thickness $1/H$ are shown.
Since the transition ceases to be first-order at the critical thickness $H_{\rm x}$
the isotropic and nematic branches terminate.
\begin{figure}
\begin{center}
\includegraphics[width=8cm]{cumulant}
\caption{\label{cumulant} Cumulant analysis of \eq{eqll} using $p=8$. Shown is
$U_1$ versus $\epsilon$ using several values of the lateral film extension $L$,
for (a) $H=1$, (b) $H=2$, and (c) $H=4$. The value of $\epsilon$ at the cumulant
intersection yields the transition inverse temperature $\epsilon_\infty(H)$ of
the thermodynamic limit $L \to \infty$.}
\end{center}
\end{figure}
\begin{table}
\caption{\label{crit} Phase transition properties for $H<H_{\rm x}$, for the
continuous IN transition. Listed is the transition inverse temperature
$\epsilon_\infty(H)$, along with the exponents $\tilde\beta$ and $\tilde\gamma$,
versus the film thickness $H$. The results refer to $p=8$ in \eq{eqll}.}
\begin{ruledtabular}
\begin{tabular}{cccc}
$H$ & $\epsilon_\infty(H)$ & $\tilde\beta$ & $\tilde\gamma$ \\ \hline
1 & 2.450 & 0.19 & 1.63 \\
2 & 1.864 & 0.17 & 1.67 \\
3 & 1.716 & 0.15 & 1.71 \\
4 & 1.650 & 0.10 & 1.81 \\
\end{tabular}
\end{ruledtabular}
\end{table}
We now consider $H < H_{\rm x}$. For $H=1$ and $p=2$ in \eq{eqll}, recent results
\cite{citeulike:5158181, citeulike:5158109} indicate the absence of any phase
transition (not even a continuous transition of the KT type). Part of the
evidence is based on the failure of the Binder cumulant to intersect. At a
continuous phase transition the ratio $U_1 = \avg{S^2} / \avg{S}^2$ becomes
independent of system size \cite{binder:1981, citeulike:6170526}, where $S$ is
the nematic order parameter. In simulations, this can be used to locate a
continuous transition, by plotting $U_1$ versus $\epsilon$ for several system
sizes $L$. At the transition inverse temperature $\epsilon_\infty(H)$ of the
film in the thermodynamic limit the curves for different lateral extensions $L$
are expected to intersect. While for $H=1$ and $p=2$ no intersections are found
\cite{citeulike:5158109}, the result for $p=8$ is radically different, see
\fig{cumulant}(a). Shown is $U_1$ versus $\epsilon$ using $H=1$ for several
values of $L$. The curves clearly intersect and so we conclude that a
continuous phase transition is taking place. This result strikingly illustrates
the non-universality of the IN transition: whether a transition occurs for $H=1$
is determined by the exponent $p$ in \eq{eqll}, i.e.~a microscopic detail! Using
$p=8$ we have verified that continuous transitions exist for all values of the
film thickness $H < H_{\rm x}$. The results for $H=2$ and $H=4$, where $H$ is
approaching $H_{\rm x}$, are shown in \fig{cumulant}(b) and~(c), both of which
reveal cumulant intersections.
The fact that the cumulants intersect is a consequence of hyperscaling. At the
transition inverse temperature $\epsilon_\infty(H)$ the order parameter decays
$\avg{S} \propto L^{-{\tilde\beta}}$, while the susceptibility $\chi = N ( \avg{S^2} -
\avg{S}^2)$ diverges $\chi \propto L^{\tilde\gamma}$. The exponents ${\tilde\beta}$ and
${\tilde\gamma}$ are connected via the hyperscaling relation
\begin{equation}\label{eqhs}
{\tilde\gamma} + 2{\tilde\beta} = d,
\end{equation}
with spatial dimension $d=2$ for a film. This relation implies that the order
parameter and its root-mean-square deviation scale $\propto L^x$ with the same
exponent~$x$. Consequently, appropriately constructed cumulant ratios, such as
$U_1$, become independent of $L$ whenever hyperscaling holds. By tuning the
inverse temperature $\epsilon$ we have determined $\epsilon_\infty(H)$ in our
simulations by requiring that the scaling of $\avg{S}$ and $\chi$ with $L$
conforms to hyperscaling, i.e.~we numerically solved \eq{eqhs}. A solution to
\eq{eqhs} for each $H<H_{\rm x}$ could indeed be found; the resulting estimates of
$\epsilon_\infty(H)$, as well as the exponents ${\tilde\beta}$ and ${\tilde\gamma}$, are
listed in \tab{crit}. As expected, $\epsilon_\infty(H)$ in \tab{crit} is close
to the cumulant intersections of \fig{cumulant}, the discrepancy being less than
0.1~\%. Note also that $\epsilon_\infty(H)$ increases with decreasing $H$. The
latter is consistent with the general tendency of confinement to lower phase
transition temperatures.
For $H=1$ the system has become 2D and the exponents reflect \ahum{pure} values,
free from any crossover effects. Note that the exponents for $H=1$ deviate
significantly from the XY~values ${\tilde\beta}_{XY}=1/8$ and ${\tilde\gamma}_{XY}=7/4$
\cite{kosterlitz:1974}, strongly suggesting a different universality class. For
$H>1$, the trend is that ${\tilde\beta} \to 0$, while ${\tilde\gamma} \to 2$. Our
interpretation is that, for $1<H<H_{\rm x}$, one observes crossover scaling behavior
\cite{citeulike:3687059}, governed by two competing fixed points: one being the
first-order transition at $H=H_{\rm x}$ and the other being the continuous transition
at $H=1$. The exponents for $1<H<H_{\rm x}$ are therefore \ahum{effective exponents},
with values between those of the $H=1$ system, and the \ahum{first-order} values
${\tilde\beta}_{\rm 1st} = 0$ and ${\tilde\gamma}_{\rm 1st} = d = 2$ \cite{citeulike:3610966}.
Note that effective exponents do not convey any fundamental information: if we
were able to simulate arbitrarily large $L$ values arbitrarily close to the
transition inverse temperature, the same exponents as for the $H=1$ system would
be found.
\begin{figure}
\begin{center}
\includegraphics[width=1.0\columnwidth]{sus}
\caption{\label{fig:sus} Variation of the susceptibility $\chi$ with inverse
temperature $\epsilon$ for several values of the lateral film extension $L$,
using film thicknesses (a) $H=1$ and (b) $H=2$. Both of these values $H<H_{\rm x}$ and
so the IN transition is continuous. Note the logarithmic vertical scale! The
data were obtained using $p=8$ in \eq{eqll}.}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=1.0\columnwidth]{cv}
\caption{\label{fig:cv} Variation of the specific heat $C$ with inverse
temperature $\epsilon$ for several values of the lateral film extension $L$,
using film thicknesses (a) $H=1$ and (b) $H=2$. Both of these values $H<H_{\rm x}$ and
so the IN transition is continuous. The data were obtained using $p=8$ in
\eq{eqll}.}
\end{center}
\end{figure}
The important result to take from this analysis is that for $p=8$ in \eq{eqll}
and $H<H_{\rm x}$ a continuous phase transition is found; by enforcing hyperscaling
the transition inverse temperature $\epsilon_\infty(H)$ can be quite accurately
obtained.
We now consider how the nematic order parameter $S$ depends on
$\epsilon$ and $L$; a typical result is shown in \fig{op}(c) where $H=2$ was
used. We note that $S$ increases with $\epsilon$ and that the slope
$dS/d\epsilon$ reaches a maximum close to $\epsilon_\infty(H)$. In contrast to
the first-order transitions observed for $H>H_{\rm x}$, $S$ does not saturate at high
$\epsilon$ but decreases steadily with increasing $L$; this behavior is
typical for all $H<H_{\rm x}$. Our simulation data thus suggest the absence of
long-range nematic order in the thermodynamic limit when $H<H_{\rm x}$. This rules out
a conventional critical point, since then the order parameter grows as a power
law $S \propto t^\beta$, $t>0$, implying $S>0$ in the nematic phase, with
distance from the transition
\begin{equation}\label{rel}
t = \epsilon - \epsilon_\infty(H)
\end{equation}
and $\beta$ the critical exponent of the order parameter. It is most likely,
therefore, that the continuous transition we observe is a topological transition
of the KT type \cite{kosterlitz.thouless:1972}.
Consistent with the KT scenario is our previous result of the order parameter
decaying $\avg{S} \propto L^{-{\tilde\beta}}$, and the susceptibility diverging $\chi
\propto L^{\tilde\gamma}$, whilst obeying hyperscaling. For completeness, we provide in
\fig{fig:sus} some raw simulation data for the susceptibility. Clearly visible
is that $\chi$ versus $\epsilon$ reveals a maximum, becoming more pronounced for
increasing $L$. In principle, the inverse temperature $\epsilon_{L,\chi}(H)$
where the susceptibility reaches its maximum, in a film of thickness $H$ and
lateral extension $L$, can be extrapolated using
\begin{equation}\label{fit}
\epsilon_{L,\chi}(H) = \epsilon_\infty(H) + \frac{b}{ \ln(L/c)^{1/\nu} },
\end{equation}
with non-universal constants $b$ and $c$, and where the exponent $\nu$
characterizes the exponential divergence of the correlation length $\xi \propto
\exp(b t^\nu)$ for $t<0$, with $t$ given by \eq{rel}. For the XY~model it holds
that $\nu_{XY}=1/2$, but since we did not recover XY~exponents in \tab{crit}
the application of \eq{fit} requires that $\nu$ be fitted also, implying a
4-parameter fit. We found that such a fitting procedure was numerically
difficult to perform, and hence we did not determine $\epsilon_\infty(H)$ in
this manner.
Finally, we note that also the specific heat, defined in \eq{eqcv}, is
consistent with the KT scenario. Plotted in \fig{fig:cv} is the variation of $C$
with $\epsilon$ for several $L$, using two values of the film thickness. In both
cases a maximum is revealed, but for $H=1$ it grows only weakly with $L$. This
is consistent with a negative specific heat exponent, implying that $C$ remains
finite in the thermodynamic limit, which agrees with the KT scenario. For $H=2$
we observe that $C$ already grows quite profoundly with $L$. We again attribute
this to the crossover to a first-order transition where, ultimately, the
specific heat maximum should scale $\propto L^{\tilde\alpha}$, with ${\tilde\alpha}_{\rm
1st}=2$, see also \fig{fig:xover}(b).
\subsection{first-order transitions}
\begin{figure}
\begin{center}
\includegraphics[width=1.0\columnwidth]{first}
\caption{\label{fig:first} Scaling analysis of \eq{eqll} using $p=20$ and film
thickness $H=1$. In (a) we show the variation of the barrier $\Delta F$ versus
$L$, while in (b) the specific heat maximum $C_{L, \rm max}(H)$ versus $L$ is
shown. Both these results indicate a first-order phase transition, even though
the system is purely 2D. The dashed line in (a) is the result of a linear fit
through the origin. The curve in (b) is a fit to the form $C_{L, \rm max}(H)
\propto L^{\tilde\alpha}$; we obtain ${\tilde\alpha} \approx 1.98$, which is very close to
${\tilde\alpha}_{\rm 1st}=2$ of a first-order phase transition in 2D.}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=1.0\columnwidth]{p20_pd}
\caption{\label{fig:p20_pd} Capillary phase diagram of \eq{eqll} using $p=20$.
Shown is the variation of the coexisting energy densities $\rho_{\rm iso}(H)$
and $\rho_{\rm nem}(H)$ versus the inverse film thickness $1/H$. In this case no
crossover occurs and the IN transition remains first-order irrespective of $H$.
The isotropic and nematic branches of the binodal therefore do not terminate,
but continue all the way to $H=1$.}
\end{center}
\end{figure}
We now consider the IN transition using $p=20$ in \eq{eqll}. In this case, the
transition is strongly first-order, even in the thin-film limit. The application
of the Lee-Kosterlitz scaling method for $H=1$ is shown in \fig{fig:first}(a),
where the linear increase of the barrier $\Delta F$ with $L$ is clearly visible.
The scaling of the specific heat maximum also confirms a first-order
transition, see \fig{fig:first}(b), showing the expected quadratic
dependence of $C_{L, \rm max}(H)$ on $L$. Since increasing the film thickness
makes the transition more strongly first-order, it is clear that for $p=20$ no
crossover can occur. In the capillary phase diagram, see \fig{fig:p20_pd}, the
isotropic and nematic branches of the coexisting energy densities do not
terminate, but continue all the way to $H \to 1$.
\subsection{Kelvin equation}
\begin{table}
\caption{\label{tab:first} Dependence of the transition inverse temperature
$\epsilon_\infty(H)$ on the film thickness $H$, for selected values of $H$ where
the IN transition is first-order. Results are shown for exponents $p=8$ and
$p=20$ in \eq{eqll}. The variation of $\epsilon_\infty(H)$ with $H$ should
follow the Kelvin equation, see \eq{eq:kelvin}. The bottom three lines list the
bulk $(H \to \infty)$ transition inverse temperature $\epsilon_\infty$ the bulk
latent heat density $\cal L_\infty$, and the bulk interfacial tension
$\gamma_\infty$, which are required in order to compare to the Kelvin
equation.}
\begin{ruledtabular}
\begin{tabular}{ccc|ccc}
$p=8$ & $H$ & $\epsilon_\infty(H)$ & $p=20$ & $H$ & $\epsilon_\infty(H)$ \\ \hline
& 5 & 1.614 & & 1 & 2.769 \\
& 6 & 1.593 & & 2 & 2.175 \\
& 7 & 1.578 & & 4 & 1.962 \\
& 8 & 1.568 & & 8 & 1.874 \\
& 10 & 1.555 & & 10 & 1.858 \\
& 15 & 1.540 & & 30 & 1.821 \\
& 30 & 1.528 & & & \\
& 50 & 1.525 & & & \\
& 100 & 1.522 & & & \\ \hline
& $\epsilon_\infty$ & 1.521 & & $\epsilon_\infty$ & 1.806 \\
& $\cal L_\infty$ & 0.909 & & $\cal L_\infty$ & 1.727 \\
& $\gamma_\infty$ & 0.06 & & $\gamma_\infty$ & 0.30 \\
\end{tabular}
\end{ruledtabular}
\end{table}
Finally, we study the variation of the inverse transition temperature $\epsilon_\infty(H)$
with the film thickness for those cases where the IN transition is first-order.
We expect $\epsilon_\infty(H)$ to fit to the Kelvin equation \cite{citeulike:3687084} as
\begin{equation}\label{eq:kelvin}
\Delta \epsilon \equiv 1 - \epsilon_\infty / \epsilon_\infty(H) =
\frac{ 2 \gamma_\infty }{ {\cal L}_\infty H},
\end{equation}
where $\gamma_\infty$ is the bulk $(H \to \infty)$ interfacial tension,
$\epsilon_\infty$ the bulk IN transition inverse temperature and $\cal
L_\infty$ the bulk latent heat density. In deriving this equation complete
wetting is assumed \cite{citeulike:3687084}. All quantities that appear in
\eq{eq:kelvin} can, in principle, be extracted from finite-size simulation data
with relative ease. For example, $\epsilon_\infty(H)$ at a first-order transition can be
obtained from $\epsilon_{L,k}(H)$; the latter is defined as the inverse
temperature where the ratio of the peak areas in the energy distribution $P(E)$
equals $k$. For an optimal value $k = k_{\rm opt}$, which can be found using
trial-and-error, the $L$-dependence in $\epsilon_{L,k}(H)$ becomes negligible
and $\epsilon_\infty(H)$ can be accurately obtained \cite{citeulike:5202797}. The resulting
estimates of the transition inverse temperatures, for both $p=8$ and $p=20$, are
provided in \tab{tab:first}, using only values of the film
thickness where the transition is first-order.
\begin{figure}
\begin{center}
\includegraphics[width=1.0\columnwidth]{stretches}
\caption{\label{fig:stretches} Plots of $\ln P$ as obtained in completely
periodic simulation boxes of size $10 \times 10 \times 30$ (solid lines) and $10
\times 10 \times 60$ (dashed lines) for (a) $p=20$ and (b) $p=8$. The height of
the peaks $\Delta F$ is related to the interfacial tension $\gamma_\infty$ via
\eq{eq:binder}. The distance $\Delta \rho$ between the peaks is a measure of the
latent heat density $\cal L_\infty$. In these plots $\epsilon$ was tuned to
yield an approximately horizontal region between the peaks.}
\end{center}
\end{figure}
Similar to previously, bulk $H \to \infty$ results are obtained using $L \times
L \times L$ systems with periodic boundaries in all directions. The bulk latent
heat density ${\cal L}_\infty$ is obtained from the specific heat maximum using
\eq{eq:cvlh} and is once again extrapolated to $L \to \infty$, where now
$N=L^3$. The resulting estimate of $\cal L_\infty$ is also listed in
\tab{tab:first}. To obtain the bulk interfacial tension $\gamma_\infty$ we use
the method of Binder \cite{binder:1982}. Simulating a large and stretched $L
\times L \times D$ system $D > L$ with periodic boundaries in all directions,
the logarithm of the energy distribution $P(E)$ reveals a pronounced flat region
between the peaks, see \fig{fig:stretches}. The flat region indicates that the
isotropic and nematic phase coexist with only small interactions between the two
interfaces. Hence, the average peak height $\Delta F$ is related to the bulk
interfacial tension
\begin{equation}\label{eq:binder}
\gamma_\infty = \lim_{L \to \infty} \gamma_L, \hspace{5mm}
\gamma_L = \Delta F / (2L^2),
\end{equation}
yielding an elegant method of obtaining~$\gamma_\infty$. Provided $L$ is large
enough, the result should not depend on the elongation $D$, but inspection of
\fig{fig:stretches} reveals this is not quite true, especially for $p=8$.
This could indicate that some interaction between the interfaces remains, or
that $L$ was not large enough. In any case, using the largest available system
size, we obtain $\gamma_\infty \approx 0.05$ for $p=8$ and $\gamma_\infty
\approx 0.29$ for $p=20$ (in units of $k_B T$ per lattice spacing squared).
Alternatively, $\gamma_L$ can be measured in a cubic periodic system of size
$L$ and extrapolation to $L \to \infty$ using
\begin{equation}\label{eq:gamma}
\gamma_L = \gamma_\infty + c_1 \ln L / L^2 + c_2 / L^2,
\end{equation}
with constants $c_i$, can be attempted \cite{binder:1982}. When using this
procedure we obtain slightly higher values of the interfacial tension, namely
$\gamma_\infty \approx 0.08$ and $\gamma_\infty \approx 0.31$ for $p=8$ and
$p=20$ respectively. Hence, for $p=20$ the estimates for $\gamma_\infty$ agree
reasonably well, whereas for $p=8$ some discrepancy remains. In \tab{tab:first}
the average of both estimates is provided.
\begin{figure}
\begin{center}
\includegraphics[width=1.0\columnwidth]{kelvin}
\caption{\label{fig:kelvin} Test of the Kelvin equation. Plotted is the inverse
temperature shift $\Delta \epsilon$ of \eq{eq:kelvin} versus the inverse film
thickness $1/H$, using exponents $p=8$ (a) and $p=20$ (b) in \eq{eqll}.}
\end{center}
\end{figure}
We now have all quantities needed to put the Kelvin equation to the test, see
\eq{eq:kelvin}. Shown in \fig{fig:kelvin} is $\Delta \epsilon$ versus $1/H$, for
both $p=8$ and $p=20$, using only values of $H$ where the transition is
first-order. Provided the Kelvin equation holds, the resulting plots should be
linear. For $p=8$ this is clearly not the case; only in the limit $1/H \to 0$,
i.e.~where the transition is strongly first-order, is agreement observed. In
contrast, using $p=20$ the Kelvin equation holds for all values of the film
thickness, including $H=1$. The slope $a$ of the lines in \fig{fig:kelvin} can
be obtained from a fit; following \eq{eq:kelvin} it is expected that $a = 2
\gamma_\infty / \cal L_\infty$, allowing for a stringent quantitative test. For
$p=20$ we obtain by fitting $a \approx 0.34$, which is in excellent agreement
with $2\gamma_\infty / {\cal L}_\infty \approx 0.35$ calculated using the
independent estimates of \tab{tab:first}. For $p=8$ the fit yields $a \approx
0.14$, where only the largest three values of $H$ were used. Once again, this
is in excellent agreement with $2\gamma_\infty / {\cal L}_\infty \approx 0.13$
obtained from \tab{tab:first}.
\section{Discussion and Summary}
\label{the_end}
In this paper we have provided new results regarding the IN transition in liquid
crystals confined between neutral walls. The main conclusion to be taken from
this work is that a single universal scenario describing the nature of this
transition as function of the film thickness $H$ does not exist. Using a
generalized version of the LL model, we have explicitly demonstrated that the
first-order IN transition can terminate at a critical thickness $H_{\rm x}$, below
which it becomes continuous, or that it can stay first-order irrespective of
$H$. The scenario that takes place is determined by a single parameter in the
Hamiltonian, namely $p$ in \eq{eqll}, which sets the \ahum{sharpness} of the
pair interaction. When the transition is sufficiently strongly first-order
excellent agreement with the Kelvin equation is also obtained. In particular, we
not only observe the $1/H$ shift of the transition inverse temperature but also
the prefactor of the shift is in quantitative agreement with the {\it
independently measured} bulk latent heat and interfacial tension. However, when
the IN transition is only weakly first-order clear deviations appear and the
Kelvin equation significantly underestimates the inverse temperature shift, see
\fig{fig:kelvin}(a).
The two different manifestations of the confined IN transition presented in this
work yield two distinct phase diagram topologies: one where the isotropic and
nematic branches of the binodal terminate at the critical thickness $H_{\rm x}$ and
one where they continue irrespective of $H$. It is of some interest to compare
the resulting phase diagrams to other works. The topology of the $p=8$ phase
diagram, see \fig{phase}, is commonly encountered in confined colloidal rods and
plates \cite{citeulike:3683414, citeulike:4922416, citeulike:3687040,
dijkstra.roij.ea:2001}. To facilitate the comparison, the energy density in
\fig{phase} should be interpreted as the analogue of the particle density in
colloidal systems. In agreement with \fig{phase}, the first-order IN transition
in colloidal systems also terminates at a critical thickness
\cite{citeulike:3683414, citeulike:4922416, citeulike:3687040,
dijkstra.roij.ea:2001}. It is also interesting to see that the nematic branch of
the binodal in \fig{phase} shows rather extreme outward curvature as the bulk
limit is approached. Colloidal platelets reveal similar behavior, albeit that
here the effect appears in the isotropic branch \cite{citeulike:3687040}. In
contrast with colloidal systems is the fact that \eq{eqll} with $p=8$ in the
bulk limit yields a first-order transition that is too strong. Defining the
relative strength of the transition as
\begin{equation}
r = \frac{ \rho_{\rm nem} - \rho_{\rm iso} }{
\rho_{\rm nem} + \rho_{\rm iso} },
\end{equation}
we obtain $r \approx 0.38$ for \eq{eqll} with $p=8$, while Onsager's exact
solution \cite{onsager:1949} for infinitely slender rods yields $r \approx
0.12$. This discrepancy can be fixed by using a lower $p$ in \eq{eqll}. For
instance, $p=5$ gives $r \approx 0.15$ \cite{citeulike:5202797}, which is much
closer to Onsager's result. Note that $p=5$ still exceeds the original LL value
$p=2$. Indeed, it has been pointed out that the original LL model yields a bulk
IN transition that is too weakly first-order compared to what is observed in
fluids of rods \cite{citeulike:4484545}.
The second phase diagram topology, where the binodal branches do not terminate
in thin films, is obtained for $p=20$ in \eq{eqll}, see \fig{fig:p20_pd}. The
resulting phase diagram is of fundamental importance, since it clearly
demonstrates that first-order IN transitions in thin films are also possible and
that the crossover to a continuous transition need not necessarily take place.
It is interesting that experiments so far have not produced clear evidence of a
continuous IN transition in thin films \cite{citeulike:4067023,
citeulike:4006159, citeulike:2811025, citeulike:3991276}. This is consistent
with a phase diagram topology as shown in \fig{fig:p20_pd}. However, it is
obvious that the model of \eq{eqll} with $p=20$ does not capture the bulk limit
correctly, since the bulk IN transition ought to be weak, whereas $p=20$ yields
a very strong first-order transition. Clearly, some features are still lacking
in \eq{eqll}, for instance a coupling between the orientational and spatial
degrees of freedom of the particles, as well as anchoring effects at the walls.
Investigations which incorporate these effects are possible directions for
future work.
Finally, using $p=8$ and $H<H_{\rm x}$ our results show that a genuine continuous IN
transition can also take place. Since long-range nematic order is not observed
a transition of the KT type \cite{kosterlitz.thouless:1972} is the most likely
scenario. This result is interesting because using $p=2$ one finds that
\eq{eqll} is without any kind of phase transition in the thin-film limit
\cite{citeulike:5158181, citeulike:5158109}. Hence, the nature of the IN
transition in thin films is ultimately determined by microscopic details. This
means that a single universality class for the IN transition cannot exist.
Depending on the details of the interaction, there can be both first-order
and continuous transitions as well as no transition occurring at all.
\acknowledgments
This work was supported by the {\it Deutsche Forschungsgemeinschaft}
under the Emmy Noether program (VI~483/1-1).
|
\section{Introduction}
\noindent Given a self-affine tiling ${\mathcal T}$ of ${\mathbb R}^d$, we consider
the tiling space, or ``hull'' $X_{{\mathcal T}}$, defined as the orbit
closure of ${\mathcal T}$ in the ``local'' topology (please see the next
section for precise definitions and statements). The translation
action by ${\mathbb R}^d$ is uniquely ergodic, so we get a
measure-preserving tiling dynamical system $(X_{{\mathcal T}},{\mathbb R}^d,\mu)$.
We are interested in its spectral properties, specifically, in the
discrete component of the spectrum which may be defined as the
closed linear span of the eigenfunctions in $L^2(X_{{\mathcal T}},\mu)$. In
particular, we would like to know when the tiling system is {\em
weakly mixing}, which means absence of non-trivial eigenfunctions.
Our results give a complete answer to these questions in terms of
the {\em expansion matrix} $\phi$ of the tiling, under the
assumption that it is diagonalizable over ${\mathbb C}$ and its
eigenvalues are algebraic
conjugates of the same multiplicity. Let
${\Lambda}=\{\lambda_1,\ldots,\lambda_{d}\}={\em Spec}(\phi)$ be the set of (real and complex)
eigenvalues of $\phi$. It is known \cite{Ken.thesis,LS} that all
$\lambda_i$ are algebraic integers. Following Mauduit \cite{Maud}, we
say that they form a {\em Pisot family} if for every $\lambda\in
{\Lambda}$ and every Galois conjugate $\lambda'$ of $\lambda$, if
$\lambda'\not\in {\Lambda}$, then $|\lambda'| < 1$. We prove that
$(X_{{\mathcal T}},{\mathbb R}^d,\mu)$ has a relatively dense set of eigenvalues
(equivalently, the set of eigenvalues of full rank $d$) if and
only if ${\Lambda}$ is a Pisot family, and this is also equivalent to
$(X_{{\mathcal T}},{\mathbb R}^d,\mu)$ being not weakly mixing.
An example shows that if the multiplicities of the eigenvalues of $\phi$ are not equal, even if
${\em Spec}(\phi)$ is a Pisot family, the set of eigenvalues of the tiling dynamical system
(not to be confused with ${\em Spec}(\phi)$) may fail to be relatively dense in ${\mathbb R}^d$.
Special cases of our theorem were established earlier: for
self-similar tilings of ${\mathbb R}^d$, with $d\le 2$, in \cite{soltil},
and for self-similar tilings of ${\mathbb R}^d$ with a pure dilation
expanding matrix $\theta I$, in \cite{sol-eigen}. The present
paper covers a much more general self-affine case.
Additional motivation for our work comes from the theory of {\em
aperiodic order} and mathematics of quasicrystals. Considering
specially chosen ``control points'' in the tiles, we obtain a
Delone set ${\mathcal C}$, that is, a uniformly discrete and relatively
dense subset of ${\mathbb R}^d$, which is a {\em substitution Delone set},
see \cite{lawa}. (We should note that in geometric analysis Delone
sets are usually called {\em separated nets}.) It can be viewed as
an atomic configuration, and it turns out that its {\em
diffraction spectrum} is, in a certain precise sense, a ``part''
of the dynamical spectrum of the system $(X_{{\mathcal T}},{\mathbb R}^d,\mu)$, with
the Bragg peaks (sharp bright spots on the diffraction picture)
coming from the eigenvalues, see \cite{Dw,LMS,Gouere,BL}. Thus,
for instance, weak mixing implies that there are no Bragg peaks,
which indicates a certain level of ``disorder.''
A Delone set $Y\subset {\mathbb R}^d$ is {\em Meyer} if it is
relatively dense and $Y-Y$ is uniformly discrete. Answering a
question of Lagarias \cite{Lag}, we showed in \cite{LS} that
having a relatively dense set of Bragg peaks is equivalent to $Y$
being Meyer, for a primitive substitution Delone set associated
with a self-affine tiling. Our results in this paper imply that
this is also characterized by the Pisot family condition (under
our assumptions on $\phi$). The notion of {\em Meyer set} proved to be important
in the study of aperiodic order, see e.g. \cite{AL, Moody, LMS2,
Lee, BLM}. In \cite{AL}, under the Meyer set assumption for a
substitution tiling, a computational algorithm is developed to
decide whether the dynamical system has pure discrete spectrum.
It should be noted that we do not address the question of
pure discrete spectrum in the present paper but focus on the
discrete spectral component.
\subsection{Structure of the proof.}
A criterion for ${\bf x}\in {\mathbb R}^d$ to be an eigenvalue of the system
$(X_{{\mathcal T}},{\mathbb R}^d,\mu)$ was obtained in \cite{soltil,sol-eigen}. From
it, the necessity of the Pisot family condition follows rather
easily, see \cite{soltil,Robi}. For the converse, we need
information on the location of control points, which is a
manifestation of certain ``rigidity'' of self-affine tilings. For
a self-similar tiling of the plane ${\mathbb C}\approx {\mathbb R}^2$ with a complex
expansion constant $\lambda$, a result of Kenyon \cite{Ken} says that
the control points are contained in ${\mathbb Z}[\lambda]{\bf a}$, for some
${\bf a}\in {\mathbb C}$. We need an extension of this statement to the
higher-dimensional self-affine case, which is a key result for us
(see Theorem~\ref{isomorphicImageOfC-contained} below). This
result does not depend on the Pisot family condition. We prove it
using the techniques developed by Thurston \cite{Thur} in the
2-dimensional self-similar case and extended to the
$d$-dimensional self-affine case in \cite{Ken.thesis,KS}, where a
necessary condition (which may be called the ``Perron family
condition'') for $\phi$ to be an expansion map was obtained. It is
here that we use the assumption that $\phi$ is diagonalizable: the
analog of the main theorem in \cite{KS} is open even for a
$2\times 2$ Jordan block.
\section{Definitions and statement of results}
\noindent We briefly review the basic definitions of tilings and
substitution tilings (see \cite{LMS2,Robi} for more details). We begin with a set of types (or colors)
$\{1,\ldots, \kappa\}$, which we fix once and for all. A {\em
tile} in ${\mathbb R}^d$ is defined as a pair $T=(A,i)$ where $A=\mbox{\rm supp}(T)$
(the support of $T$) is a compact set in ${\mathbb R}^d$ which is the
closure of its interior, and $i=l(T)\in \{1,\ldots, \kappa\}$ is
the type of $T$. We let $g+T = (g+A,i)$ for $g\in {\mathbb R}^d$. We say
that a set $P$ of tiles is a {\em patch} if the number of tiles in
$P$ is finite and the tiles of $P$ have mutually disjoint
interiors. A tiling of ${\mathbb R}^d$ is a set ${\mathcal T}$ of tiles such that
${\mathbb R}^d = \cup \{\mbox{\rm supp}(T) : T \in {\mathcal T}\}$ and distinct tiles have
disjoint interiors. Given a tiling ${\mathcal T}$, finite sets of tiles of
${\mathcal T}$ are called ${\mathcal T}$-patches. For $A \subset {\mathbb R}^d$, let
$[A]^{{\mathcal T}} = \{T \in {\mathcal T} : \mbox{\rm supp}(T) \cap A \neq \emptyset\}$.
We always assume that any two ${\mathcal T}$-tiles with the same color are
translationally equivalent. (Hence there are finitely many
${\mathcal T}$-tiles up to translation.)
We say that a tiling ${\mathcal T}$ has {\em finite local complexity (FLC)}
if for each radius $R > 0$ there are only finitely many
translational classes of patches whose support lies in some ball
of radius $R$.
A tiling ${\mathcal T}$ is said to be {\em repetitive} if translations of any given patch occur uniformly dense in
${\mathbb R}^d$; more precisely, for any ${\mathcal T}$-patch $P$, there exists $R>0$ such that every ball of radius $R$ contains
a translated copy of $P$.
Given a tiling ${\mathcal T}$, we define the {\em tiling space} as the
orbit closure of ${\mathcal T}$ under the translation action: $X_{{\mathcal T}} =
\overline{\{-g + {\mathcal T}:\,g\in {\mathbb R}^d\}}$, in the well-known ``local
topology'': for a small ${\mbox{$\epsilon$}}>0$ two tilings ${\mathcal S}_1,{\mathcal S}_2$ are
${\mbox{$\epsilon$}}$-close if ${\mathcal S}_1$ and ${\mathcal S}_2$ agree on the ball of radius
${\mbox{$\epsilon$}}^{-1}$ around the origin, after a translation of size less
than ${\mbox{$\epsilon$}}$. It is known that $X_{\mathcal T}$ is compact whenever ${\mathcal T}$
has FLC. Thus we get a topological dynamical system
$(X_{{\mathcal T}},{\mathbb R}^d)$ where ${\mathbb R}^d$ acts by translations. This system is
minimal (i.e.\ every orbit is dense) whenever ${\mathcal T}$ is repetitive.
Let $\mu$ be an invariant Borel probability measure for the
action; then we get a measure-preserving system $(X_{{\mathcal T}}, {\mathbb R}^d,
\mu)$. Such a measure always exists; under the natural assumption
of {\em uniform patch frequencies}, it is unique, see \cite{LMS}.
Tiling dynamical system have been investigated in a large number of papers; we do not provide an exhaustive bibliography, but mention a few: \cite{Petersen,CS,HRS,Kel}.
They have also been studied as translation surfaces or ${\mathbb R}^d$-solenoids \cite{BG,Gamb}.
\begin{defi}
{\em A vector $\mbox{\boldmath{$\alpha$}} =(\alpha_1,\ldots,\alpha_d) \in {\mathbb R}^d$ is
said to be an {\em eigenvalue} for the ${\mathbb R}^d$-action if there
exists an eigenfunction $f\in L^2(X_{{\mathcal T}},\mu)$, that is, $\
f\not\equiv 0$ and for all $g\in {\mathbb R}^d$ and $\mu$-almost all
${\mathcal S}\in X_{{\mathcal T}}$,
\begin{equation} \label{def-eig1}
f({\mathcal S}-g) = e^{2 \pi i \langle g, \alpha\rangle} f({\mathcal S}).
\end{equation}
Here $\langle \cdot,\cdot \rangle$ denotes the standard scalar
product in ${\mathbb R}^d$.}
\end{defi}
Note that this ``eigenvalue'' is actually a vector. In physics it might be
called a ``wave vector.'' We can also speak about eigenvalues for
the topological dynamical system $(X_{{\mathcal T}},{\mathbb R}^d)$; then the eigenfunction should be in
$C(X_{{\mathcal T}})$ and the equation (\ref{def-eig1}) should hold everywhere.
Next we define substitution tilings. Let $\phi$ be an expanding
linear map in ${\mathbb R}^d$, which means that all its eigenvalues are
greater than one in modulus.
The following definition is
essentially due to Thurston \cite{Thur}.
\begin{defi}\label{def-subst}
{\em Let ${\mathcal A} = \{T_1,\ldots,T_{\kappa}\}$ be a finite set of
tiles in ${\mathbb R}^d$ such that $T_i=(A_i,i)$; we will call them {\em
prototiles}. Denote by ${\mathcal A}^+$ the set of patches made of tiles
each of which is a translate of one of $T_i$'s. We say that
$\omega: {\mathcal A} \to {\mathcal A}^+$ is a {\em tile-substitution} (or simply
{\em substitution}) with expansion map $\phi$ if there exist
finite sets ${\mathcal D}_{ij}\subset {\mathbb R}^d$ for $i,j \le \kappa$, such that
\begin{equation}\label{subdiv}
\omega(T_j)= \{u+T_i:\ u\in {\mathcal D}_{ij},\ i=1,\ldots, \kappa\} \ \ \
\mbox{for} \ j\le \kappa,
\end{equation}
with
$$
\phi A_j = \bigcup_{i=1}^{\kappa} ({\mathcal D}_{ij}+A_i).
$$
Here all sets in the right-hand side must have disjoint interiors;
it is possible for some of the ${\mathcal D}_{ij}$ to be empty.}
\end{defi}
The substitution (\ref{subdiv}) is extended to all translates of
prototiles by $\omega(x+T_j)= \phi x + \omega(T_j)$, and to patches and
tilings by $\omega(P)=\cup\{\omega(T):\ T\in P\}$. The substitution
$\omega$ can be iterated, producing larger and larger patches
$\omega^k(T_j)$. To the substitution $\omega$ we associate its $\kappa
\times \kappa$ substitution matrix with the entries $\sharp
({\mathcal D}_{ij})$. The substitution $\omega$ is called {\em primitive} if
the substitution matrix is primitive. We say that ${\mathcal T}$ is a fixed
point of a substitution if $\omega({\mathcal T}) = {\mathcal T}$.
\begin{defi} \label{def-saf}
{\em A repetitive fixed point of a primitive tile-substitution
with FLC is called a {\em self-affine tiling}. It is called {\em
self-similar} if the expansion map is a similitude, that is,
$|\phi(x)| = {\theta} |x|$ for all $x\in {\mathbb R}^d$, with some ${\theta}>1$.}
\end{defi}
\begin{remark}
{\em 1. A fixed point of a primitive tile-substitution is not
necessarily of finite local complexity, see \cite{Danzer,FraRob}.
Thus we have to assume FLC explicitly.
2. It is well-known (and easy to see, e.g.\ in the one-dimensional
case) that the fixed point may be repetitive even for a
non-primitive substitution. Conversely, the fixed point of a
primitive substitution need not be repetitive. (However, if the
tile-substitution is primitive and the fixed point tiling ${\mathcal T}$
has a tile which contains the origin in its interior, then ${\mathcal T}$
is repetitive \cite{Prag}.)
3. For a self-similar tiling of ${\mathbb R}^d$, with $d\le 2$, we can
speak of an expansion factor; it is a real number if $d=1$ and a
complex number if $d=2$ (we then view the plane as a complex
plane).}
\end{remark}
An important question, first raised by Thurston \cite{Thur}, is to
characterize which expanding linear maps may occur as expansion
maps for self-affine (self-similar) tilings. It is pointed out in
\cite{Thur} that in one dimension, $\lambda$ is an expansion factor
if and only if ${\theta}=|\lambda|$ is a {\em Perron number}, that is, an
algebraic integer greater than one whose Galois conjugates are all
strictly less than ${\theta}$ in modulus (necessity follows from the
Perron-Frobenius theorem and sufficiency follows from a result of
Lind \cite{Lind}). In two dimensions, Thurston \cite{Thur} proved
that if $\lambda$ is a complex expansion factor of a self-similar
tiling, then $\lambda$ is a {\em complex Perron number}, that is, an
algebraic integer whose Galois conjugates, other than $\overline{\lambda}$,
are all less than $|\lambda|$ in modulus. The following theorem was
stated in \cite{Ken.thesis}, but complete proof was not available
until recently.
\begin{theorem} \cite{Ken.thesis,KS} \label{th-KS}
Let $\phi$ be a diagonalizable (over ${\mathbb C}$) expansion map on
${\mathbb R}^d$, and let ${\mathcal T}$ be a self-affine tiling of ${\mathbb R}^d$ with
expansion $\phi$. Then
{\bf (i)} every eigenvalue of $\phi$ is an algebraic integer;
{\bf (ii)}
if $\lambda$ is an eigenvalue of $\phi$ of multiplicity $k$ and ${\gamma}$ is
an algebraic conjugate of $\lambda$, then either $|{\gamma}| < |\lambda|$, or ${\gamma}$
is also an eigenvalue of $\phi$ of multiplicity greater or equal to $k$.
\end{theorem}
\begin{remark}
{\em 1. Note that if $|{\gamma}|=|\lambda|$ in part (ii) of the theorem,
then the multiplicities of ${\gamma}$ and $\lambda$ are the same.
2. It is conjectured that the condition on $\phi$ in the theorem
is also sufficient. There are partial results in this direction
\cite{Kenyon.construction}; see \cite{KS} for a discussion.}
\end{remark}
For a self-affine tiling ${\mathcal T}$, the
corresponding tiling dynamical system $(X_{{\mathcal T}}, {\mathbb R}^d)$ is
uniquely ergodic, see \cite{LMS2, Robi}. Denote by $\mu$ the
unique invariant probability measure.
There is a rich structure associated with self-affine tiling dynamical systems.
As a side remark, we mention that the substitution map $\omega$ extends to an endomorphism of the tiling space,
which is hyperbolic in a certain sense, see \cite{AP}.
The partition of the tiling space according to the type of the tile containing the origin provides a Markov partition for $\omega$. The situation is especially nice
when ${\mathcal T}$ is non-periodic, which is equivalent to $\omega$ being invertible \cite{solucp}. In order to state our results we need the following.
\begin{defi} \cite{Maud}
{\em A set ${\Lambda}$ of algebraic integers is called a {\em Pisot
family} if for every $\lambda\in {\Lambda}$, if ${\gamma}$ is an algebraic
conjugate of $\lambda$ and ${\gamma}\not\in {\Lambda}$, then $|{\gamma}|< 1$.
Otherwise ${\Lambda}$ is called {\em non-Pisot}.}
\end{defi}
In this paper we assume that:
\begin{itemize}
\item all the eigenvalues of $\phi$ are algebraic conjugates
with the same multiplicity.
\end{itemize}
Let $Spec(\phi)$ be the set of all eigenvalues of $\phi$ (the spectrum of $\phi$).
By assumption, there exists a monic irreducible polynomial
$p(t) \in {\mathbb Z}[t]$ (the minimal polynomial) such that $p(\lambda) = 0$
for all $\lambda \in Spec(\phi)$.
\begin{theorem} \label{th-main}
Let ${\mathcal T}$ be a self-affine tiling of ${\mathbb R}^d$ with a diagonalizable
expansion map $\phi$. Suppose that all the eigenvalues of $\phi$
are algebraic conjugates with the same multiplicity. Then
the following are equivalent:\\
{\bf (i)} The set of eigenvalues of $(X_{{\mathcal T}},{\mathbb R}^d,\mu)$ is
relatively dense in ${\mathbb R}^d$.\\
{\bf (ii)} $Spec(\phi)$ is a Pisot family. \\
{\bf (iii)} The system $(X_{{\mathcal T}},{\mathbb R}^d,\mu)$ is not weakly
mixing (i.e., it has eigenvalues other than ${\bf 0}$).
\end{theorem}
\begin{remark}
{\em 1. In part (i) we could equally well talk about the
topological dynamical system $(X_{{\mathcal T}},{\mathbb R}^d)$ since every
eigenfunction may be chosen to be continuous \cite{sol-eigen}.
2. The necessity of the Pisot family condition for self-affine tiling systems that are not weakly mixing
was proved by Robinson \cite{Robi} in a more
general case; it is a consequence of \cite{soltil}.}
\end{remark}
\begin{example} {\em (i) In Fig.\,2 and Fig.\,3 of \cite{KS} a self-affine tiling ${\mathcal T}_1$ is given, with the diagonal expansion matrix ${\rm Diag}[\lambda_1,\lambda_2]$ where
$\lambda_1\approx 2.19869$ and $\lambda_2\approx -1.91223$ are roots of the polynomial $x^3-x^2-4x+3$. Observe that $\{\lambda_1,\lambda_2\}$ is a Pisot family, hence
the set of eigenvalues for the associated dynamical system is relatively dense in ${\mathbb R}^2$.
(ii) The assumption of equal multiplicity cannot be dropped from Theorem~\ref{th-main}.
Indeed, consider the tiling ${\mathcal T}$ which is a ``direct product'' of ${\mathcal T}_1$ defined in (i) and
a self-similar tiling ${\mathcal T}_2$ of ${\mathbb R}$ with expansion $\lambda_1$.
Such a tiling ${\mathcal T}_2$ exists by \cite{Lind} (see \cite{Solnotes} for more details)
since $\lambda_1$ is a Perron number. Direct product substitution tilings have been studied by S. Mozes \cite{mozes} and N. P. Frank \cite{natalie}. It is
easy to see that the set of eigenvalues for the dynamical system $(X_{\Tk},{\mathbb R}^3)$ is obtained as a direct sum of those which
correspond to the systems $(X_{{\mathcal T}_1},{\mathbb R}^2)$ and $(X_{{\mathcal T}_2},{\mathbb R})$. By \cite{soltil}, the system $(X_{{\mathcal T}_2},{\mathbb R})$ is
weakly mixing, because $\lambda_1$ is not a Pisot number. Thus, the tiling ${\mathcal T}$ has expansion map
$\phi={\rm Diag}[\lambda_1,\lambda_2,\lambda_1]$ for which ${\em Spec}(\phi)$ is a Pisot family, but the associated dynamical
system does not have a relatively dense set of eigenvalues. }
\end{example}
Next we state our result on Meyer sets.
Recall that a Delone set is a relatively dense and uniformly
discrete subset of ${\mathbb R}^d$.
\begin{defi}
{\em A Delone set $Y$ is called a {\em Meyer set} if
$Y-Y$ is uniformly discrete.}
\end{defi}
There is a standard way to choose distinguished points in the tiles of a self-affine tiling so that they form a $\phi$-invariant Delone set. They are called
{\em control points}.
\begin{defi} \cite{Thur,Prag}
{\em Let ${\mathcal T}$ be a fixed point of a primitive substitution with
expansion map $\phi$. For each ${\mathcal T}$-tile $T$, fix a tile $\gamma
T$ in the patch $\omega (T)$; choose $\gamma T$ with the same
relative position for all tiles of the same type. This defines a
map $\gamma : {\mathcal T} \to {\mathcal T}$ called the {\em tile map}. Then define
the {\em control point} for a tile $T \in {\mathcal T}$ by
\[ \{c(T)\} = \bigcap_{n=0}^{\infty} \phi^{-n}(\gamma^n T).\]
}
\end{defi}
\noindent The control points have the following properties:
\begin{itemize}
\item[(a)] $T' = T + c(T') - c(T)$, for any tiles $T, T'$ of the
same type; \item[(b)] $\phi(c(T)) = c(\gamma T)$, for $T \in {\mathcal T}$.
\end{itemize}
Control points are also fixed for tiles of any tiling $\mathcal{S}
\in X_{{\mathcal T}}$: they have the same relative position as in
${\mathcal T}$-tiles. Note that the choice of control points is non-unique, but there are only finitely many possibilities, determined by the choice of the tile map.
Let
\[ {\mathcal C}:= {\mathcal C}({\mathcal T}) = \{ c(T) : T \in {\mathcal T} \}\]
be a set of control points of the tiling ${\mathcal T}$ in ${\mathbb R}^d$. Let
$$
\Xi := \Xi({\mathcal T}) = \bigcup_{i=1}^{\kappa} ({\mathcal C}_i - {\mathcal C}_i),
$$
where ${\mathcal C}_i$ is the set of control points of tiles of type $i$.
Equivalently, $\Xi$ is the set of translation vectors between two
${\mathcal T}$-tiles of the same type.
\begin{cor} \label{cor-Meyer}
Let ${\mathcal T}$ be a self-affine tiling of ${\mathbb R}^d$ with a diagonalizable
expansion map $\phi$. Suppose that all the eigenvalues of $\phi$
are algebraic conjugates with the same multiplicity. Then the set
of control points ${\mathcal C}$ is Meyer if and only if $Spec(\phi)$ is a
Pisot family.
\end{cor}
This is an immediate consequence of Theorem~\ref{th-main} and
\cite[Th.\,4.14]{LS}.
\section{Preliminaries}
\noindent Recall that $\phi$ is assumed to be diagonalizable over
${\mathbb C}$. For a complex eigenvalue $\lambda$ of $\phi$, the $2 \times 2$
diagonal block $\left[
\begin{array}{ll}
\lambda & 0 \\
0 & \overline{\lambda}
\end{array} \right] $ is similar to a real $2 \times 2$ matrix
\begin{eqnarray} \label{eq-mult}
\left[ \begin{array}{rr}
a & -b \\
b & a
\end{array} \right] = S^{-1}\left[\begin{array}{ll}
\lambda & 0 \\
0 & \overline{\lambda}
\end{array} \right]S,
\end{eqnarray} where $\lambda = a + ib$, $a, b \in
{\mathbb R}$, and $S = \frac{1}{\sqrt{2}}\left[\begin{array}{rr} 1 & i \\ 1 & -i \end{array} \right]$.
Since $\phi$ is diagonalizable
over ${\mathbb C}$, we can assume, by appropriate choice of basis, that
$\phi$ is in the real canonical form of the linear map, see
\cite[Th.\,6.4.2]{HS}. This means that $\phi$ is block-diagonal,
with the diagonal entries equal to $\lambda$ corresponding to real
eigenvalues, and diagonal $2 \times 2$ blocks of the form $\left[
\begin{array}{rr}
a_j & -b_j \\
b_j & a_j
\end{array} \right]$ corresponding to complex eigenvalues
$a_j + i b_j$.
Let $J$ be the multiplicity of each eigenvalue of $\phi$. We can
write
\[ \phi = \left[
\begin{array}{ccc}
\psi_1 & \cdots & {O} \\
\vdots & \ddots & \vdots \\
{O} & \cdots & \psi_J
\end{array} \right] \ \mbox{and} \ \
\psi_j = \psi := \left[
\begin{array}{ccc}
A_{1} & \cdots & 0 \\
\vdots & \ddots & \vdots \\
0 & \cdots & A_{s+t}
\end{array} \right] \ \ \ \mbox{for any $1 \le j \le J$} \]
where $A_{k}$ is a real $1\times 1$ matrix for $1 \le k \le s$, a
real $2\times 2$ matrix of the form $\left[
\begin{array}{lr}
a_{k} & -b_{k} \\
b_{k} & a_{k}
\end{array} \right] $
for $s + 1 \le k \le s + t$, and ${O}$ is the $(s+2t) \times (s+2t)$
zero matrix. Then the eigenvalues of $\psi$ are
$$\lambda_{1}, \ldots, \lambda_{s},\lambda_{s+1},\overline{\lambda}_{s+1},\ldots,\lambda_{s + t},\overline{\lambda}_{s+t}.$$
Let $m: = s +2t$; this is the size of the matrix $\psi$.
For each $1 \le j \le J$, let
\begin{equation}
H_{j} = \{0\}^{(j-1)m} \times {\mathbb R}^{m} \times
\{0\}^{d- jm}\,. \nonumber
\end{equation}
Further, for each $H_{j}$ we have the direct sum decomposition
$$
H_j = \bigoplus_{k=1}^{s + t} E_{jk},
$$
such that each $E_{jk}$ is $\phi,\phi^{-1}$-invariant and
$\phi|_{E_{jk}} \approx A_{k}$, identifying $E_{jk}$ with ${\mathbb R}$ or
${\mathbb R}^2$. Define a norm on ${\mathbb R}^d$ by
\begin{equation} \label{def-norm}
\|{\bf x}\| = \max_{j,k}\|{\bf x}_{jk}\|\ \ \mbox{for}\
{\bf x}= \sum_{j=1}^J \sum_{k=1}^{s+t} {\bf x}_{jk},\ {\bf x}_{jk} \in E_{jk},
\end{equation}
where $\|{\bf x}_{jk}\|$ is the Euclidean norm on $E_{jk}$, so that
$\|\phi {\bf x}_{jk}\| = |\lambda_{jk}| \|{\bf x}_{jk}\|$. Let
$$
{\mathbb Q}[\phi]:= \{p(\phi):\ p \in {\mathbb Q}[x]\},\ \ {\mathbb Z}[\phi]:= \{p(\phi):\ p \in {\mathbb Z}[x]\}.
$$
Let $P_{j}$ be the canonical projection of ${\mathbb R}^d$ onto $H_{j}$
such that \begin{eqnarray} \label{def-projection} P_{j}({\bf x}) = {\bf x}_{j}, \end{eqnarray}
where ${\bf x} = {\bf x}_{1} + \cdots + {\bf x}_{J}$ and ${\bf x}_{j} \in H_{j}$
with $1 \le j \le J$. Let $\phi_{j} = \phi|_{H_{j}}$.
\medskip
We define $\mbox{\boldmath{$\alpha$}}_{j} \in H_{j}$ such that for each $1 \le n \le
d$, \begin{eqnarray} \label{def-alpha} (\mbox{\boldmath{$\alpha$}}_{j})_{n} =
\left\{\begin{array}{ll}
1 \ \ \ & \mbox{if} \ \ (j-1)m + 1 \le n \le jm;\\
0 \ \ \ & \mbox{else} \,.
\end{array} \right.
\end{eqnarray}
The next theorem is a key result of the paper; it is the
manifestation of rigidity alluded in the Introduction.
\begin{theorem} \label{isomorphicImageOfC-contained}
Let ${\mathcal T}$ be a self-affine tiling of ${\mathbb R}^d$ with a diagonalizable
expansion map $\phi$. Suppose that all the eigenvalues of $\phi$
are algebraic conjugates with the same multiplicity. Then there
exists an isomorphism $\rho: {\mathbb R}^d \to {\mathbb R}^d$ such that
\[ \mbox{$\rho \phi =
\phi \rho$ \ \ \ and} \ \ \ \mathcal{C} \subset \rho({\mathbb Z}[\phi] \mbox{\boldmath{$\alpha$}}_{1} + \cdots +
{\mathbb Z}[\phi] \mbox{\boldmath{$\alpha$}}_{J})\,,\] where $\mbox{\boldmath{$\alpha$}}_j$, $1 \le j \le J$, are given as above.
\end{theorem}
The reason we call this ``rigidity'' is by analogy with \cite[Th.\,9]{Ken} (see the discussion at the beginning of the proof in \cite{Ken}).
We give a proof of Theorem~\ref{isomorphicImageOfC-contained} in Section \ref{Structure of
control point set} below and make use of it in proving the main
theorem in Section \ref{Meyer-property}. Note that the choice of
$\mbox{\boldmath{$\alpha$}}_j$ is rather arbitrary; it is ``hidden'' in the linear
isomorphism $\rho$.
Now we continue with the preliminaries; we
need to handle the real and complex eigenvalues a little bit
differently.
Consider the linear injective map ${\mathcal F}:{\mathbb R}^m\,\to {\mathbb R}^s\oplus
{\mathbb C}^{2t}$ given by \begin{eqnarray} \label{def-Fk} {\mathcal F}(x_1,\ldots, x_s,
x_{s+1},\ldots,x_{s+2t}) =\end{eqnarray}$$ =\left(x_1,\ldots, x_s,
\frac{x_{s+1}+ix_{s+2}}{\sqrt{2}},
\frac{x_{s+1}-ix_{s+2}}{\sqrt{2}},\ldots, \frac{x_{s+2t-1} +
ix_{s+2t}}{\sqrt{2}},
\frac{x_{s+2t-1}-ix_{s+2t}}{\sqrt{2}}\right)\,. $$ In other words,
identifying $H_j$ with ${\mathbb R}^m$, we apply the transformation $S$
from (\ref{eq-mult}) in every subspace $E_{jk}$,
$k=s+1,\ldots,s+t$. In view of (\ref{eq-mult}), we have \begin{eqnarray}
\label{eq-mult2} {\mathcal F}(\psi {\bf x}) = D {\mathcal F}({\bf x})\ \ \ \mbox{and}\ \ \
{\mathcal F}(\psi^T {\bf x}) = \overline{D} {\mathcal F}({\bf x}), \end{eqnarray} where \begin{eqnarray} \label{def-diag}
D={\rm Diag}[\lambda_1,\ldots,\lambda_s,\lambda_{s+1},\overline{\lambda}_{s+1}\ldots,
\lambda_{s+t},\overline{\lambda}_{s+t}]\end{eqnarray} is a diagonal matrix.
\medskip
The following lemma is well-known and easy to prove using the
Vandermonde matrix.
\begin{lemma} \label{independent-phi-alphas-I}
Let $D$ be a diagonal matrix on ${\mathbb C}^m$ with distinct complex
eigenvalues. Let ${\bf z} = [z_1, \cdots, z_m ]^T \in {\mathbb C}^m$ be such
that $z_k \neq 0$ for all $1 \le k \le m$. Then $\{{\bf z}, D {\bf z},
\cdots, D^{m-1} {\bf z} \}$ is linearly independent over ${\mathbb C}$.
\end{lemma}
\begin{cor} \label{independent-phi-alphas}
Suppose that ${\bf x} \in {\mathbb R}^m$ is such that ${\bf z}={\mathcal F}({\bf x})$ has all
$(m)$ non-zero coordinates. Then both $\{{\bf x},
\psi{\bf x},\ldots,\psi^{m-1}{\bf x}\}$ and $\{{\bf x}, \psi^T {\bf x},\ldots,
(\psi^T)^{m-1}{\bf x}\}$ are linearly independent over ${\mathbb R}$.
\end{cor}
\begin{sloppypar}
\begin{proof}
We have
$$
{\mathcal F}(\{{\bf x},\ldots,\psi^{m-1}{\bf x}\}) = \{{\bf z},\ldots,D^{m-1}{\bf z}\}
$$
by (\ref{eq-mult2}). By Lemma~\ref{independent-phi-alphas-I}, the
set $\{{\bf z},\ldots,D^{m-1}{\bf z}\}$ is independent over ${\mathbb C}$ and hence
$\{{\bf x}, \psi{\bf x},\ldots,\psi^{m-1}{\bf x}\}$ is independent over ${\mathbb R}$,
using the fact that ${\mathcal F}$ is injective. The proof for the second set (with
transpose matrices) is exactly the same. \end{proof}
\end{sloppypar}
\begin{cor} \label{basisWithAlphas}
The set $W := \{\mbox{\boldmath{$\alpha$}}_{1}, \ldots, \phi^{m-1} \mbox{\boldmath{$\alpha$}}_{1},
\ldots, \mbox{\boldmath{$\alpha$}}_{J}, \ldots, \phi^{m-1}\mbox{\boldmath{$\alpha$}}_{J}\}$ forms a
basis of ${\mathbb R}^d$.
\end{cor}
\begin{proof}
Identifying $H_j$ with ${\mathbb R}^m$, we have $\phi_j =\phi|_{H_j}\approx
\psi$ and use the isomorphism ${\mathcal F}$ defined above. In view of
(\ref{def-alpha}), all the components of
${\bf z}_j={\mathcal F}(\mbox{\boldmath{$\alpha$}}_j)$ are non-zero, so the claim follows from
Corollary~\ref{independent-phi-alphas}.
\end{proof}
For ${\bf x},{\bf y}\in {\mathbb R}^m$ we use the standard scalar product $\langle
{\bf x},{\bf y} \rangle = \sum_{k=1}^m x_k y_k$, and for ${\bf z},{\bf u} \in
{\mathbb R}^{s} \oplus {\mathbb C}^{2t}$ the scalar product is given by
$$
\langle {\bf z},{\bf u} \rangle_{_{\mathbb C}} = \sum_{k=1}^{s+2t} z_k \overline{u}_k.
$$
Observe that \begin{eqnarray} \label{eq-scalar} \langle {\bf x},{\bf y} \rangle =
\langle {\mathcal F}({\bf x}),{\mathcal F}({\bf y}) \rangle_{_{\mathbb C}}\ \ \ \mbox{for all}\ \
{\bf x},{\bf y}\in {\mathbb R}^m. \end{eqnarray} Recall also that for any $m \times m$ matrix
$A$,
$$
\langle A{\bf x},{\bf y} \rangle = \langle {\bf x},A^T{\bf y} \rangle\ \ \ \mbox{for all}\ \ {\bf x},{\bf y}\in {\mathbb R}^m.
$$
\section{Proof of the main theorem (proof of Theorem\,\ref{th-main})} \label{Meyer-property}
\noindent Here we deduce Theorem\,\ref{th-main} from
Theorem\,\ref{isomorphicImageOfC-contained}. Recall that a set of
algebraic integers $\Theta = \{\theta_1, \cdots, \theta_r \}$ is a
{\em Pisot family} if for any $1 \le j \le r$, every Galois
conjugate $\gamma$ of $\theta_j$ with $|\gamma| \ge 1$ is
contained in $\Theta$. We denote by $\mbox{\rm dist}(x,{\mathbb Z})$ the distance from a real number $x$ to the nearest integer.
\begin{prop} \label{set-of-eigenvalues}
Let ${\mathcal T}$ be a self-affine tiling of ${\mathbb R}^d$ with a diagonalizable
expansion map $\phi$. Suppose that all the eigenvalues of $\phi$
are algebraic conjugates with the same multiplicity. If
$Spec(\phi)$ is a Pisot family, then the set of eigenvalues of
$(X_{{\mathcal T}}, {\mathbb R}^d, \mu)$ is relatively dense.
\end{prop}
\begin{proof} Recall that $\Xi = \{{\bf x} \in {\mathbb R}^d :\ \exists\ T \in {\mathcal T},\ T+{\bf x}
\in {\mathcal T}\}$ is the set of ``return vectors'' for the tiling ${\mathcal T}$,
and let $\mathcal{K} = \{{\bf x} \in {\mathbb R}^d :\ {\mathcal T} - {\bf x} = {\mathcal T} \}$ be
the set of translational periods. Clearly, ${\mathcal K} \subset \Xi
\subset {\mathcal C}-{\mathcal C}$. We know from \cite[Th.\,3.13]{sol-eigen} that
$\gamma$ is an eigenvalue for $(X_{{\mathcal T}}, {\mathbb R}^d, \mu)$ if and only
if \[ \lim_{n \to \infty} e^{2 \pi i \langle \phi^n {\bf x},
\gamma \rangle} = 1 \ \ \ \mbox{for all ${\bf x} \in \Xi$ \;\;
and}\]
\[e^{2 \pi i \langle {\bf x},
\gamma \rangle} = 1 \ \ \ \mbox{for all ${\bf x} \in
\mathcal{K}$}.\]
Let $\mbox{\boldmath{$\alpha$}}_{j}\in H_{j}$ be the vectors from (\ref{def-alpha}).
Consider them as vectors in ${\mathbb R}^m$, and let ${\mathcal F}$ be the linear
map ${\mathbb R}^m \to {\mathbb R}^s\oplus {\mathbb C}^{2t}$ given by (\ref{def-Fk}). Recall
that $\phi_j=\phi|_{H_j}$ has $s$ real and $2t$ complex
eigenvalues, and $m = s+2t$. Define $\mbox{\boldmath{$\beta$}}_{j}\in H_{j}\approx
{\mathbb R}^m$ so that
$$
({\mathcal F}(\mbox{\boldmath{$\beta$}}_{j}))_k = (\overline{{\mathcal F}(\mbox{\boldmath{$\alpha$}}_{j})})_k^{-1}\ \ \mbox{for}\ 1 \le k \le m.
$$
More explicitly,
$$
(\mbox{\boldmath{$\beta$}}_j)_k = (\mbox{\boldmath{$\alpha$}}_j)_k^{-1}\ \ \mbox{for}\ 1 \le k \le s
$$
and $$ (\mbox{\boldmath{$\beta$}}_j)_{s+2k-1} \pm i(\mbox{\boldmath{$\beta$}}_j)_{s+2k} =
\frac{2}{(\mbox{\boldmath{$\alpha$}}_j)_{s+2k-1} \mp i(\mbox{\boldmath{$\alpha$}}_j)_{s+2k}}\ \
\mbox{for}\ 1 \le k \le t.
$$
\smallskip
\noindent
Note that $\mbox{\boldmath{$\beta$}}_{j}\in H_{j}$ are well-defined, and
${\mathcal F}(\mbox{\boldmath{$\beta$}}_{j})$ have all non-zero coordinates in $H_{j}$. Thus,
\begin{eqnarray} \label{prebasis} B_{j} := \{\mbox{\boldmath{$\beta$}}_{j}, \ldots,
(\phi^T)^{m-1}\mbox{\boldmath{$\beta$}}_{j}\} \end{eqnarray} is a basis of $H_{j}$ by
Corollary~\ref{independent-phi-alphas} (note that $H_j$ is also
$\phi^T$-invariant and $\phi^T|_{H_j}$ is isomorphic to $\psi^T$).
It follows that $B:=\bigcup_{j=1}^J B_j$ is a basis of ${\mathbb R}^d$. We
will show that all elements of the set $(\rho^T)^{-1}(\phi^T)^K B$ are
eigenvalues for the tiling dynamical system, for $K$ sufficiently
large.
By the definition of $\mbox{\boldmath{$\beta$}}_{j}$, in view of (\ref{eq-scalar})
and (\ref{eq-mult2}), for any $n \in {\mathbb Z}_{\ge 0}$ and $0 \le l <
m$,
\begin{eqnarray}
\langle \phi^n \mbox{\boldmath{$\alpha$}}_{j}, (\phi^T)^l \mbox{\boldmath{$\beta$}}_{j} \rangle & = & \langle \phi^{n+l}\mbox{\boldmath{$\alpha$}}_j, \mbox{\boldmath{$\beta$}}_j\rangle \nonumber \\
& = & \langle {\mathcal F}(\phi^{n+l}\mbox{\boldmath{$\alpha$}}_j), {\mathcal F}(\mbox{\boldmath{$\beta$}}_i)\rangle_{\mathbb C} \nonumber \\
& = & \langle D^{n+l} {\mathcal F}(\mbox{\boldmath{$\alpha$}}_i), {\mathcal F}(\mbox{\boldmath{$\beta$}}_i)\rangle_{_{\mathbb C}} \nonumber \\
& = & \sum_{k=1}^s \lambda_k^{n+l} + \sum_{k=s+1}^{s+t} (\lambda_k^{n+l}+\overline{\lambda}_k^{n+l}). \label{eq-PV}
\end{eqnarray}
Here $D$ is the diagonal matrix from (\ref{def-diag}).
Since
$Spec(\phi)$ is a Pisot family, it follows
that $\mbox{\rm dist}(\langle \phi^n \mbox{\boldmath{$\alpha$}}_{j},
(\phi^T)^l \mbox{\boldmath{$\beta$}}_{j} \rangle,{\mathbb Z})\to 0$, as $n\to \infty$. (This is a standard argument: the sum of $(n+l)$-th powers of all
zeros of a polynomial in ${\mathbb Z}[x]$ is an integer, hence the distance from the sum in (\ref{eq-PV}) to ${\mathbb Z}$ is
bounded by the sum of the moduli of $(n+l)$-th powers of their remaining conjugates, which are all less than one in modulus.
Thus, this distance tends to zero exponentially fast.)
Observe also that $\langle
\phi^n \mbox{\boldmath{$\alpha$}}_{u}, \mbox{\boldmath{$\beta$}}_{j}\rangle = 0$ if $u\ne j$, hence
\[ \lim_{n \to \infty} e^{2 \pi i \langle \phi^n {\bf y}, (\phi^T)^l {\mbox{\tiny\boldmath${\beta}$}}_{j} \rangle} = 1
\ \ \ \mbox{for all} \ {\bf y} \in {\mathbb Z}[\phi]\mbox{\boldmath{$\alpha$}}_{1} + \cdots +
{\mathbb Z}[\phi]\mbox{\boldmath{$\alpha$}}_{J} \,. \] Therefore, by
Theorem~\ref{isomorphicImageOfC-contained}, using that $\Xi
\subset {\mathcal C} - {\mathcal C}$, we obtain \begin{eqnarray}
\label{first-eigenvalue-condition} \lim_{n \to \infty} e^{2 \pi i
\langle \phi^n {\bf x}, (\rho^T)^{-1}(\phi^T)^l {\mbox{\tiny\boldmath${\beta}$}}_{j} \rangle}
= 1 \ \ \ \mbox{for all} \ {\bf x} \in \Xi \,. \end{eqnarray} Furthermore, by
\cite[Cor.\,4.4]{sol-eigen}, the convergence is uniform in ${\bf x}
\in \Xi$, that is,
\[\lim_{n \to \infty} \sup_{{\bf x} \in \Xi} |e^{2\pi i \langle \phi^n {\bf x}, (\rho^T)^{-1}(\phi^T)^l {\mbox{\tiny\boldmath${\beta}$}}_{j} \rangle} - 1| = 0 \,.\]
Recall that $\mathcal{K} \subset \Xi$, and $\mathcal{K}$ is a
discrete subgroup in ${\mathbb R}^d$. So for every ${\bf x} \in {\mathcal K}$, $$
\lim_{n \to \infty} \sup_{\stackrel{k \in {\mathbb Z}_+}{{\bf x} \in
\mathcal{K}}} |e^{2\pi i \langle \phi^n (k {\bf x}),
(\rho^T)^{-1}(\phi^T)^l {\mbox{\tiny\boldmath${\beta}$}}_{j} \rangle} - 1| = \lim_{n \to
\infty} \sup_{\stackrel{k \in {\mathbb Z}_+}{{\bf x} \in \mathcal{K}}} |e^{2\pi
i k \langle \phi^n {\bf x}, (\rho^T)^{-1}(\phi^T)^l {\mbox{\tiny\boldmath${\beta}$}}_{j}
\rangle} - 1| = 0. $$ It follows that there exists $K_{l} \in
{\mathbb Z}_+$ such that for any $n \ge K_{l}$, for all ${\bf x}\in {\mathcal K}$, \begin{eqnarray}
\label{supremumOfexp-eigenvalue} \sup_{\stackrel{k \in {\mathbb Z}_+}{{\bf x}
\in \mathcal{K}}} |e^{2\pi i k \langle \phi^n {\bf x},
(\rho^T)^{-1}(\phi^T)^l {\mbox{\tiny\boldmath${\beta}$}}_{j} \rangle} - 1| < 1/2 \,. \end{eqnarray}
However, unless $\langle \phi^n {\bf x}, (\rho^T)^{-1}(\phi^T)^l
\mbox{\boldmath{$\beta$}}_{j} \rangle \in {\mathbb Z}$ for all ${\bf x} \in \mathcal{K}$,
(\ref{supremumOfexp-eigenvalue}) does not hold. Thus
\[e^{2 \pi i \langle \phi^n {\bf x}, (\rho^T)^{-1}(\phi^T)^l \mbox{\tiny\boldmath${\beta}$}_{j} \rangle} = e^{2 \pi i \langle {\bf x}, (\rho^T)^{-1}(\phi^T)^{n+l} \mbox{\tiny\boldmath${\beta}$}_{j} \rangle} = 1 \ \
\mbox{for all ${\bf x} \in \mathcal{K}$ and all $n \ge K_{l}$}\,.\]
Let $ K = {\rm max} \{K_{l}:\ 0 \le l < m \}. $ Then \begin{eqnarray}
\label{second-eigenvalue-condition} e^{2 \pi i \langle {\bf x},
(\rho^T)^{-1}(\phi^T)^{K+l} \mbox{\tiny\boldmath${\beta}$}_{j} \rangle} = 1 \ \ \mbox{for
all ${\bf x} \in \mathcal{K}$}\,. \end{eqnarray} So from
(\ref{first-eigenvalue-condition}) and
(\ref{second-eigenvalue-condition}) it follows that
$(\rho^T)^{-1}(\phi^T)^{K+l} \mbox{\boldmath{$\beta$}}_{j}$ is an eigenvalue of
$(X_{{\mathcal T}}, {\mathbb R}^d, \mu)$ for $l = 0,\ldots,m-1$. We have shown that
all vectors of the set $(\rho^T)^{-1}(\phi^T)^K B$, where $B =
\bigcup_j B_j$ and $B_j$ are given by (\ref{prebasis}), are
eigenvalues of $(X_{{\mathcal T}}, {\mathbb R}^d, \mu)$. We know $\phi^T$ is
invertible (it is expanding), $\rho$ is a linear isomorphism, and
$B$ is a basis of ${\mathbb R}^d$, hence we obtain a basis of ${\mathbb R}^d$
consisting of eigenvalues. Integer linear combinations of
eigenvalues are eigenvalues as well, so the set of eigenvalues of
$(X_{{\mathcal T}}, {\mathbb R}^d, \mu)$ is relatively dense in ${\mathbb R}^d$.
\end{proof}
The next lemma is essentially due to Robinson \cite{Robi} in a
more general case; we provide a proof for completeness.
\begin{lemma} \label{eigenvalueImplyPisotFamily}
If $\mbox{\boldmath{$\gamma$}}$ is a non-zero eigenvalue of $(X_{{\mathcal T}}, {\mathbb R}^d, \mu)$,
then $Spec(\phi)$ is a Pisot family.
\end{lemma}
\begin{proof}
Let ${\bf x} \in \Xi$. By Theorem~\ref{isomorphicImageOfC-contained}
we have ${\bf x} = \rho(\sum_{j=1}^J p_j(\phi)\mbox{\boldmath{$\alpha$}}_j)$ for some
polynomials $p_j \in {\mathbb Z}[x]$. Let $(\rho^T \mbox{\boldmath{$\gamma$}})_j = p_j(\rho^T
\mbox{\boldmath{$\gamma$}})$. We again use the linear injective map ${\mathcal F}:\,H_j\approx
{\mathbb R}^m \to {\mathbb R}^s \oplus {\mathbb C}^{2t}$ defined by (\ref{def-Fk}) and
obtain, using (\ref{eq-scalar}) and (\ref{eq-mult2}),
\begin{eqnarray}
\langle \phi^n{\bf x},\mbox{\boldmath{$\gamma$}} \rangle & = & \sum_{j=1}^J \langle \phi^n p_{j}(\phi) \mbox{\boldmath{$\alpha$}}_{j}, (\rho^T \mbox{\boldmath{$\gamma$}})_{j} \rangle \nonumber \\
& = & \sum_{j=1}^J \langle {\mathcal F}(\phi^n p_{j}(\phi)\mbox{\boldmath{$\alpha$}}_{j}), {\mathcal F}((\rho^T\mbox{\boldmath{$\gamma$}})_{j}) \rangle \nonumber \\
& = & \sum_{j=1}^J \left(\sum_{k=1}^s \lambda_k^n p_j(\lambda_k) z_{jk}\zeta_{jk} + \sum_{k=s+1}^{s+t} 2{\rm Re}[\lambda_k^n p_j(\lambda_k) z_{j,2k-s-1}\zeta_{j,2k-s-1}]\right)\nonumber \\
& = & \sum_{k=1}^s c_k \lambda_k^n + \sum_{k=s+1}^{s+t} (c_k \lambda_k^n + \overline{c}_k \overline{\lambda}_k^n), \nonumber
\end{eqnarray}
where $(z_{jk})_{k=1}^{s+2t} = {\mathcal F}(\mbox{\boldmath{$\alpha$}}_j)$,
$(\zeta_{jk})_{k=1}^{s+2t} = {\mathcal F}((\rho^T \mbox{\boldmath{$\gamma$}})_j)$, and $c_k$
are some complex numbers. By the assumption that $\mbox{\boldmath{$\gamma$}}$ is
an eigenvalue and \cite[Th.\,4.3]{soltil} we have \begin{eqnarray}
\label{dunduk} \mbox{\rm dist}\left(\sum_{k=1}^s c_k \lambda_k^n + \sum_{k=s+1}^{s+t}
(c_k \lambda_k^n + \overline{c}_k \overline{\lambda}_k^n),\ {\mathbb Z}\right) = \mbox{\rm dist}( \langle \phi^n {\bf x},
\mbox{\boldmath{$\gamma$}} \rangle ,{\mathbb Z}) \stackrel{n \to \infty}{\longrightarrow} 0. \end{eqnarray}
Since $\Xi$ is relatively dense in ${\mathbb R}^d$ and $\mbox{\boldmath{$\gamma$}}\ne {\bf 0}$, we
can easily make sure that $\langle {\bf x},\mbox{\boldmath{$\gamma$}} \rangle\ne 0$, and
hence not all coefficients $c_k$ in (\ref{dunduk}) are equal to
zero. Then we can apply a theorem of Vijayaraghavan \cite[Th.\,4]{Vij} (a generalization of the classical result of Pisot and Vijayaraghavan)
to conclude that
${\em Spec}(\phi)$ is a Pisot family. (More precisely, we obtain that a
subset of ${\em Spec}(\phi)$ is a Pisot family, but since all elements
of ${\em Spec}(\phi)$ are conjugates and have modulus greater than one,
we get the claim.)
\end{proof}
\begin{theorem} \label{PisotFamily-relDenseEigenvalues}
Let ${\mathcal T}$ be a self-affine tiling of ${\mathbb R}^d$ with a diagonalizable
expansion map $\phi$. Suppose that all the eigenvalues of $\phi$
are algebraic conjugates with the same multiplicity. Then the
following are equivalent:
\begin{itemize}
\item[(i)] $Spec(\phi)$ forms a Pisot family;
\item[(ii)] the set of eigenvalues of $(X_{{\mathcal T}}, {\mathbb R}^d, \mu)$
is relatively dense;
\item[(iii)] $(X_{{\mathcal T}}, {\mathbb R}^d, \mu)$ is not weakly mixing;
\item[(iv)] $\mathcal{C} = \{c(T) : T \in {\mathcal T} \}$ is a Meyer
set.
\end{itemize}
\end{theorem}
\begin{proof}
(i) $\Rightarrow$ (ii) by Prop.\,\ref{set-of-eigenvalues}. \\
(ii) $\Rightarrow$ (iii) is obvious.\\
(iii) $\Rightarrow$ (i) by
Lemma~\ref{eigenvalueImplyPisotFamily}.\\
(ii) $\Leftrightarrow$ (iv) by \cite[Th.\,4.14]{LS}.
\end{proof}
Theorem\,\ref{th-main} is contained in
Theorem~\ref{PisotFamily-relDenseEigenvalues}, so it is proved as
well.
\section{Structure of the control point set (proof of Theorem\,\ref{isomorphicImageOfC-contained})}
\label{Structure of control point set}
\noindent Now we make an isomorphic transformation $\tau$ of the
tiling ${\mathcal T}$ into another tiling whose control point set contains
$\mbox{\boldmath{$\alpha$}}_{1}, \ldots, \mbox{\boldmath{$\alpha$}}_{J}$ such that $\tau$ commutes
with $\phi$. This gives the structure of the control point set of ${\mathcal T}$
that we use in proving the main theorem in Section
\ref{Meyer-property}.
\medskip
In Corollary~\ref{basisWithAlphas} we showed that $W=\{\mbox{\boldmath{$\alpha$}}_{1}, \ldots, \phi^{m-1} \mbox{\boldmath{$\alpha$}}_{1},
\ldots, \mbox{\boldmath{$\alpha$}}_{J}, \ldots, \phi^{m-1}\mbox{\boldmath{$\alpha$}}_{J}\}$ is a basis for ${\mathbb R}^d$.
Since $\mathcal{C}$ is relatively dense in ${\mathbb R}^d$, for any
$\epsilon > 0$, for
every $j=1,\ldots,J$, there exists ${\bf y}_j\in {\mathcal C}$ such that
\[ \left\| \frac{{\bf y}_j}{\|{\bf y}_j\|} - \frac{\mbox{\boldmath{$\alpha$}}_j}{\|\mbox{\boldmath{$\alpha$}}_j\|} \right\| < {\mbox{$\epsilon$}}. \]
For ${\mbox{$\epsilon$}}>0$ sufficiently small, $(j-1)m +1, \dots, (jm)$-th entries of ${\bf y}_{j}$ are
all non-zero for any $1 \le j \le J$, and the set
\[ Y := \{{\bf y}_1, \dots,
\phi^{m-1}{\bf y}_1, \dots, {\bf y}_J, \dots, \phi^{m-1}{\bf y}_J\} \subset
\mathcal{C}\] is a basis of ${\mathbb R}^d$.
We fix such a basis $Y$.
\begin{lemma} \label{isomorphism-changing-basis}
Let $\tau : {\mathbb R}^d \to {\mathbb R}^d$ be a linear map such that for each $1
\le j \le J$ and $0 \le k < m$,
\[\tau(\phi^k {\bf y}_j) = \phi^k \mbox{\boldmath{$\alpha$}}_j \,.\]
Then $\tau$ is an isomorphism of ${\mathbb R}^d$ such that $\tau \phi =
\phi \tau$.
\end{lemma}
\begin{proof}
We first notice that $\tau$ is an isomorphism of ${\mathbb R}^d$, since $Y$
and $W$ are bases of ${\mathbb R}^d$.
In order to show that $\phi\tau({\bf x}) = \tau\phi({\bf x}),\ {\bf x}\in {\mathbb R}^d$, it is enough to check this on the basis $Y$. For the vectors $\phi^k {\bf y}_j$,\ $0 \le k < m-1$, this
holds by definition, so we only need to consider $\phi^{m-1}{\bf y}_j$.
Let $p(t)$ be the characteristic polynomial of $\psi$. Then $p(\psi)=0$ by Cayley-Hamilton, and also $p(\phi)=0$, since $\phi$ is a
direct sum of $J$ copies of $\psi$. Thus
\begin{equation} \label{CH}
\phi^m =a_0I+\cdots + a_{m-1}\phi^{m-1}
\end{equation}
for some $a_0,\ldots,a_{m-1} \in {\mathbb R}$, hence
\begin{eqnarray*}
\phi\tau(\phi^{m-1}{\bf y}_j) = \phi^m \mbox{\boldmath{$\alpha$}}_j & = & a_0 \mbox{\boldmath{$\alpha$}}_j + \cdots + a_{m-1}\phi^{m-1}\mbox{\boldmath{$\alpha$}}_j \\
& = & a_0 \tau({\bf y}_j)+\cdots + a_{m-1}\tau(\phi^{m-1}{\bf y}_j) = \tau(\phi^m {\bf y}_j),
\end{eqnarray*}
as desired.
\end{proof}
Let \begin{eqnarray} \label{new-substitution-tiling} \tau({\mathcal T}):= \{\tau(T):\,
\tau(T)=(\tau(A), i), \ \mbox{where $T \in {\mathcal T}$ and $T=(A, i)$}
\}.\end{eqnarray} Note that $\tau({\mathcal T})$ is a fixed point of a primitive
substitution with the expansion map $\phi$. Indeed, $\omega'(\tau({\mathcal T}))=\tau({\mathcal T})$ where $\omega'$ is defined by
\[ \omega'(\tau(T_j)) = \{u + \tau(T_i): u \in \tau(\mathcal{D}_{ij}), i = 1, \dots, \kappa \} \ \ \
\mbox{for $j \le \kappa$,}\]
where
\[\phi\tau(A_j) = \bigcup_{i=1}^{\kappa} (\tau(\mathcal{D}_{ij}) + \tau(A_i)).\]
We define the tile map $\gamma': \tau({\mathcal T}) \to \tau({\mathcal T})$ so that
for each ${\mathcal T}$-tile $T$,
\[\gamma'(\tau(T)) = \tau(\gamma(T)).\]
We define the control point for a tile $\tau(T) \in \tau({\mathcal T})$ by
\[\{c(\tau(T))\} = \bigcap_{n=0}^{\infty} \phi^{-n}((\gamma')^n \tau({\mathcal T})).\]
Then \[c(\tau(T)) = \tau c(T).\]
Applying the isomorphism $\tau$ commuting with $\phi$, we can reduce our problem to the case when
the control point set of the tiling contains $\mbox{\boldmath{$\alpha$}}_{1}, \ldots, \mbox{\boldmath{$\alpha$}}_{J}$. Thus, in
the rest of this section (except the last paragraph which proves
Theorem\,\ref{isomorphicImageOfC-contained}), we assume that
$\mathcal{C}$ contains $\mbox{\boldmath{$\alpha$}}_{1}, \ldots, \mbox{\boldmath{$\alpha$}}_{J}$.
\medskip
The following two propositions were obtained in
\cite{KS} in a special case. They are needed to get the structure of control
point set which we use in Section \ref{Meyer-property}. In the
appendix, we provide the proof, which is similar to that in \cite{KS}, for completeness.
In the next two propositions we {\bf do not} assume that all the eigenvalues of $\phi$ are conjugates and have the same multiplicity.
Let $G_\lambda$ be the real $\phi$-invariant subspace of ${\mathbb R}^d$ corresponding to an eigenvalue $\lambda \in {\em Spec}(\phi)$.
\begin{prop} \label{prop-KenSol}
Let ${\mathcal C}$ be a set of control points for a self-affine tiling
${\mathcal T}$ of ${\mathbb R}^d$ with an expansion map $\phi:\,{\mathbb R}^d \to {\mathbb R}^d$ which
is diagonalizable over ${\mathbb C}$. Let $\mathcal{C}_{\infty} =
\bigcup_{k=0}^{\infty} \phi^{-k} \mathcal{C}$ and let $\mathcal{D}$ be
a finitely generated ${\mathbb Q}[\phi]$-module containing
$\mathcal{C}_{\infty}$. Let $H$ be a vector space over ${\mathbb R}$ and
$A:\, H \to H$ be an expanding linear map, diagonalizable over ${\mathbb C}$.
Let $g: {\mathcal D} \to H$ be such that $g = A^{-1} \circ g \circ \phi$
and
\[g(y_1) - g(y_2) = g(y_1 - y_2) \ \ \ \mbox{for any $y_1, y_2 \in {\mathcal D}$}.\]
Let $f:= g|_{{\mathcal C}_\infty}:\,{\mathcal C}_\infty \to H$ be such that
for any $y_1, y_2 \in \mathcal{C}$,
\begin{eqnarray} \label{Lipschitz-linear-on-controlPointSet}
||f(y_1) - f(y_2)|| \le C ||y_1 - y_2 || \ \ \ \mbox{for some $C > 0$.}
\end{eqnarray}
Moreover, suppose that there exist ${\gamma}>1$ and a norm
$\|\cdot\|$ in $H$ such that
\begin{equation} \label{eq-newk1}
\|A y\| \ge {\gamma}\|y\|\ \ \mbox{for any}\ y\in H.
\end{equation}
Then the following hold:
{\bf (i)} The map $f$ is uniformly continuous on ${\mathcal C}_{\infty}$,
and hence extends by continuity to a map $f:\, {\mathbb R}^d\to H$
satisfying $f\circ \phi = A\circ f$.
{\bf (ii)} For any $\lambda\in Spec(\phi)$ such that $|\lambda| =
\gamma$, and any $a\in {\mathbb R}^d$,
\[ f|_{a + G_{\lambda}} \ \mbox{is affine linear}. \]
\hfill $\square$
\end{prop}
Let $P_{\lambda}$ be the canonical projection of ${\mathbb R}^d$ to $G_{\lambda}$
commuting with $\phi$, which exists by the diagonalizability
assumption on $\phi$. Denote by $G_{\lambda}^\perp = (I -
P_{\lambda}){\mathbb R}^d$ the complementary $\phi$-invariant subspace. We
consider the set $(I - P_{\lambda})\Xi$, that is, the projection of
$\Xi$ to $G_{\lambda}^\perp$ (recall that $\Xi$ is the set of translation vectors between two ${\mathcal T}$-tiles of the same type).
In some directions this projection may look like a lattice, i.e.\ be discrete.
We consider the directions in which this set is not discrete, and denote the span of
these directions by $G'$. We will prove that $f$ is affine linear on
all $G'$ slices.
More precisely, for any ${\mbox{$\epsilon$}} > 0$, define
\begin{eqnarray} \label{def-Geps}
G_{{\mbox{$\epsilon$}}} := \mbox{Span}_{{\mathbb R}} \left(B_{{\mbox{$\epsilon$}}} \cap (I - P_{\lambda})\Xi \right)\ \ \ \mbox{and}\ \ \ G':= \bigcap_{{\mbox{$\epsilon$}} > 0} G_{{\mbox{$\epsilon$}}}.
\end{eqnarray}
Now let
\begin{eqnarray} \label{E-equal-to-E'-Elambda}
G := G' + G_{\lambda}\,.
\end{eqnarray}
Note that $G$ is a subspace of ${\mathbb R}^d$ which is $\phi$-invariant,
because \[ \mbox{$\phi \Xi \subset \Xi \ \ $ and $\phi P_\lambda = P_\lambda \phi.$}\]
\begin{prop} \label{prop-KenSol2}
Under the assumptions of Proposition~\ref{prop-KenSol}, $f|_{a + G}$ is affine linear for
any $a \in {\mathbb R}^d$. \hfill $\square$
\end{prop}
\begin{lemma} \label{G-contain-all-eigenspaces}
Let all eigenvalues of $\phi$ be algebraic conjugates with the
same multiplicity. If $\lambda$ is the smallest in modulus
eigenvalue of $\phi$, then
\[G = G'+ G_{\lambda} = {\mathbb R}^d.\]
\end{lemma}
\begin{proof} This is proved in \cite{KS} (although not stated there
explicitly). Indeed, in the last part of \cite{KS}, labeled {\em
Conclusion of the proof of Theorem 3.1}, it is proved that the
subspace $G$ (denoted $E$ there) contains, for each conjugate of
$\lambda$ greater or equal than $\lambda$ in modulus, an eigenspace of
dimension at least $\dim(G_\lambda)$. Note that in \cite{KS} the
setting is more general, of an arbitrary diagonalizable over ${\mathbb C}$
matrix $\phi$. In our case all eigenvalues are conjugates of the
same multiplicity, and $\lambda$ is the smallest in modulus, hence
$G$ contains the entire ${\mathbb R}^d$.
\end{proof}
Since ${\mathcal T}$ has FLC, the ${\mathbb Z}$-module generated by ${\mathcal C}$, denoted
by $\langle \mathcal{C} \rangle_{{\mathbb Z}}$, is finitely generated. Let
$\{{\bf v}_1, \dots, {\bf v}_N\}$ be a generating set in $\mathcal{C}$.
For each ${\bf v}_n$ with $1 \le n \le N$,
\[{\bf v}_n = a_{11}^{(n)}
\mbox{\boldmath{$\alpha$}}_1 + \cdots + a_{1m}^{(n)} \phi^{m -1} \mbox{\boldmath{$\alpha$}}_1 + \cdots +
a_{J1}^{(n)} \mbox{\boldmath{$\alpha$}}_J + \cdots + a_{Jm}^{(n)} \phi^{m -1}
\mbox{\boldmath{$\alpha$}}_J\] where $a_{jk}^{(n)} \in {\mathbb R}$, $1 \le j \le J$, and $1
\le k \le m$. Thus
\[\mathcal{C} \subset \mathcal{D}:=\sum_{j=1}^J \sum_{k=1}^{m}
\sum_{n=1}^N {\mathbb Q}[\phi] (a_{jk}^{(n)} \mbox{\boldmath{$\alpha$}}_j)= \sum_{j=1}^J \sum_{k=1}^{m}
\sum_{n=1}^N {\mathbb Q}[\phi_j] (a_{jk}^{(n)} \mbox{\boldmath{$\alpha$}}_j).\]
\begin{lemma} \label{lem-field}
${\mathbb Q}[\phi_{j}]$ is a field.
\end{lemma}
\begin{proof}
This is clearly a ring, so we just need to show that $p(\phi_j)$
has an inverse for $p \in {\mathbb Q}[x]$, if it is a non-zero matrix.
We need to use that all the eigenvalues of $\phi_j$ are conjugates
so they have the same irreducible polynomial $p(x)$. If $q(x) \in
{\mathbb Z}[x]$ is monic, such that $p(x)$ does not divide $q(x)$, then we
can find monic polynomials $h_1(x),h_2(x) \in {\mathbb Z}[x]$ such that
$h_1(x) p(x) + h_2(x) q(x) = 1$, which means that $h_2(\lambda)$ is
the inverse of $q(\lambda)$ for any eigenvalue $\lambda$ of $\phi_j$.
\end{proof}
Let $\mathcal{D}_{j} = P_{j}(\mathcal{D})$, where $P_j$ is the
canonical projection of ${\mathbb R}^d$ onto $H_j$ as defined in
(\ref{def-projection}). Observe that $\mathcal{D}_{j}$ is a vector
space over the field ${\mathbb Q}[\phi_{j}]$, so we can write
\[\mathcal{D}_{j} = \bigoplus_{t=1}^{r_j} {\mathbb Q}[\phi_j] (a_{jt} \mbox{\boldmath{$\alpha$}}_j)
= \bigoplus_{t=1}^{r_{j}} {\mathbb Q}[\phi] (a_{jt} \mbox{\boldmath{$\alpha$}}_j) ,\]
where $a_{j1} =1$, $a_{jt} \in {\mathbb R}$ with $1 \le t \le r_{j}$, and
$\{a_{j1}, \dots, a_{jr_{j}}\}$ is linearly independent over ${\mathbb Q}$.
Note that $$\mathcal{D} = \mathcal{D}_{1} \bigoplus \cdots
\bigoplus \mathcal{D}_{J}.$$
We define ${\mathbb Q}[\phi]$-module homomorphisms
\[ \sigma_{j} : {\mathcal D} \rightarrow {\mathbb Q}[\phi] \mbox{\boldmath{$\alpha$}}_{j}\]
such that \begin{eqnarray*} \left\{ \begin{array}{ll}
\sigma_{j}(a_{jt} \phi^n \mbox{\boldmath{$\alpha$}}_{j}) = \phi^n \mbox{\boldmath{$\alpha$}}_{j}
& \mbox{for any $ 1 \le t \le r_{j}$, $n \in {\mathbb Z}_{\ge 0}$}\\
\sigma_{j}(a_{uw} \phi^n \mbox{\boldmath{$\alpha$}}_{u}) ={\bf 0}
& \mbox{for any $u \neq j$, $n \in
{\mathbb Z}_{\ge 0}$}\,.
\end{array}
\right. \end{eqnarray*}
Recall that $\mathcal{C}_{\infty} := \bigcup_{k=0}^{\infty} \phi^{-k}
\mathcal{C}$. Observe that ${\mathcal D} \supset {\mathcal C}_\infty$, since $\phi^{-1}$ is a rational linear combination of $\{I,\phi,\ldots,\phi^{m-1}\}$ by (\ref{CH}).
We define $\sigma_{j}':\mathcal{C}_{\infty} \to
{\mathbb Q}[\phi] \mbox{\boldmath{$\alpha$}}_{j}$ to be the restriction of ${\sigma}_j$, that is, ${\sigma}_j': = {\sigma}_j|_{{\mathcal C}_\infty}$.
\medskip
Using the same arguments as in \cite[Lemma 5.3]{sol-eigen} (which followed \cite{Thur}), we
obtain the next lemma.
\begin{lemma} \label{inequality-on-controlPoints}
For any $\xi, \xi' \in {\mathcal C}$,
\[\|\sigma'_{j}(\xi) - \sigma'_{j}(\xi')\| \le C ||\xi - \xi'|| \ \ \ \mbox{for some $C > 0$.}\]
\end{lemma}
\medskip
Now we use Prop.\,\ref{prop-KenSol},
Prop.\,\ref{prop-KenSol2} and
Lemma\,\ref{G-contain-all-eigenspaces} to prove
Theorem\,\ref{isomorphicImageOfC-contained}, and assume that all the assumptions of the latter hold. In addition, suppose that
the set of control points contains $\mbox{\boldmath{$\alpha$}}_1,\ldots,\mbox{\boldmath{$\alpha$}}_J$.
Fix $1 \le j \le
J$. We consider the maps $g = {\sigma}_j:\,{\mathcal D}\to H_j$ and
$f = {\sigma}'_{j}:\,{\mathcal C}_\infty \to H_{j}$, and let $A = \phi_{j} =
\phi|_{H_{j}}$.
Note
that (\ref{eq-newk1}) holds with ${\gamma}$ equal to the smallest absolute value of
eigenvalues of $\phi_{j}$ (or $\phi$) and the norm defined as in (\ref{def-norm}).
Thus, all the hypotheses of Prop.\,\ref{prop-KenSol}, Prop.\,\ref{prop-KenSol2} and
Lemma\,\ref{G-contain-all-eigenspaces} are satisfied, and we obtain that
for each $1 \le j \le J$, the (extended) map $\sigma'_{j}$ is linear on
${\mathbb R}^d$ and commutes with $\phi$.
\begin{lemma} \label{rho'-is-an-isomorphism}
The map $\rho':= \sigma'_{1} + \cdots + \sigma'_{J}$ is the identity map on
${\mathbb R}^d$.
\end{lemma}
\begin{proof} Note that
\[\rho' : {\mathbb R}^d \rightarrow {\mathbb R}^d \ \ \mbox{is linear and $\rho' \phi = \phi \rho'$}
\]
and for any $1 \le j \le J$,
\begin{eqnarray*} \rho'(\mbox{\boldmath{$\alpha$}}_j) &=& (\sigma'_{1} + \cdots +
\sigma'_{J})(\mbox{\boldmath{$\alpha$}}_j) = 0 + \cdots + 0+ \sigma'_j (\mbox{\boldmath{$\alpha$}}_j) + 0+ \cdots + 0= \mbox{\boldmath{$\alpha$}}_j \,.
\end{eqnarray*}
Since $W = \{\mbox{\boldmath{$\alpha$}}_1, \dots, \phi^{m-1} \mbox{\boldmath{$\alpha$}}_1, \dots,
\mbox{\boldmath{$\alpha$}}_J, \dots, \phi^{m -1} \mbox{\boldmath{$\alpha$}}_J \}$ is a basis of ${\mathbb R}^d$,
$\rho'$ is the identity map on ${\mathbb R}^d$.
\end{proof}
Now we do not assume that the control point set of ${\mathcal T}$ contains
$\mbox{\boldmath{$\alpha$}}_{1}, \ldots, \mbox{\boldmath{$\alpha$}}_{J}$ in order to prove
Theorem~\ref{isomorphicImageOfC-contained}. Instead, we apply the above
propositions and lemmas to $\tau(\mathcal{C})$.
\medskip
\noindent {\it{Proof of
Theorem\,\ref{isomorphicImageOfC-contained}}.} By Lemma~\ref{rho'-is-an-isomorphism}, for each
$\xi \in \mathcal{C}$,
\begin{eqnarray} \label{containment-of-C}
\tau(\mathcal{\xi}) = \rho'(\tau(\mathcal{\xi})) &=& (\sigma'_{1} + \cdots +
\sigma'_{J})(\tau(\xi)) = (\sigma_{1} + \cdots +
\sigma_{J})(\tau(\xi)) \nonumber \\
& \in & {\mathbb Q}[\phi]\mbox{\boldmath{$\alpha$}}_{1} + \cdots + {\mathbb Q}[\phi]\mbox{\boldmath{$\alpha$}}_{J}\,.
\end{eqnarray}
Since $\mathcal{C}$ is finitely generated, we multiply (\ref{containment-of-C}) by a common
denominator $b \in {\mathbb Z}_+$ to get \begin{eqnarray} b
\cdot \tau(\mathcal{C}) \subset {\mathbb Z}[\phi] \mbox{\boldmath{$\alpha$}}_{1} + \cdots +
{\mathbb Z}[\phi] \mbox{\boldmath{$\alpha$}}_{J} \,. \nonumber \end{eqnarray} Let $\rho := \frac{1}{b}
\cdot \tau^{-1}$. Then
\[\mathcal{C} \subset \rho({\mathbb Z}[\phi] \mbox{\boldmath{$\alpha$}}_{1} + \cdots +
{\mathbb Z}[\phi] \mbox{\boldmath{$\alpha$}}_{J})\] where $\rho$ is an isomorphism of ${\mathbb R}^d$ which
commutes with $\phi$. \hfill $\square$
\section{Appendix}
\noindent We give the proofs of Prop.\,\ref{prop-KenSol} and
Prop.\,\ref{prop-KenSol2} after a sequence of auxiliary lemmas.
The arguments are similar to those in \cite{KS}, but we
present them in a more general form for our purposes.
Denote by $B_{R}(a)$ the open ball of radius $R$ centered at $a$ and let $B_R:= B_R(0)$. We will also write $\overline{B}_R(a)$ for the closure of $B_R(a)$.
Let $r = r({\mathcal T})
>0$ be such that for every $a\in {\mathbb R}^d$ the ball $\overline{B}_r(a)$ is covered by a tile containing $a$ (which need not be unique) and its neighbors.
Let $\lambda_{\max}$ be the largest eigenvalue of $\phi$.
\begin{lemma}
The function $f$ is uniformly continuous on $\mathcal{C}_{\infty}$.
\end{lemma}
\begin{proof}
This is very similar to \cite[Lem.\,3.4]{KS}.
It is enough to show that
\begin{eqnarray} \label{Holder-continuity} \xi_1, \xi_2 \in {\mathcal C}_\infty,\ \ \|\xi_1 - \xi_2\|\le r\ \ \Longrightarrow\ \ \|f(\xi_1) - f(\xi_2)\|\le L \|\xi_1-\xi_2\|^\alpha,
\end{eqnarray}
for $\alpha =
\frac{\log\gamma}{\log|\lambda_{\max}|}$ and some $L
> 0$ (that is, $f$ is H\"{o}lder continuous on ${\mathcal C}_\infty$).
Let $\xi_1, \xi_2 \in \mathcal{C}_{\infty}$ satisfy $||\xi_1 -
\xi_2|| = \delta \le r$. Then there exist $y_1, y_2 \in
\mathcal{C}$ such that $\phi^{-s} y_1 = \xi_1$ and $\phi^{-s} y_2
= \xi_2$ for some $s \in {\mathbb Z}_{\ge 0}$. We choose the smallest
$l \in {\mathbb Z}_{\ge 0}$ such that \[ \phi^s B_{\delta}(\phi^{-s} y_1)
\subset \phi^l B_r(\phi^{-l} y_1),\] which is equivalent to $\phi^{s-l} B_{\delta}(\phi^{-s} y_1) \subset B_r(\phi^{-l} y_1)$.
Since $\delta \le r$, we have $l \le s$ and hence $l$ is the smallest integer satisfying
$$
|\lambda_{\max}|^{s-l} \delta \le r.
$$
Thus, \begin{eqnarray} \label{randdelta-inequality} |\lambda_{\max}|^{s-l} >
\frac{r}{\delta} |\lambda_{\max}|^{-1}. \end{eqnarray} Observe that $y_2 \in
\phi^s \overline{B}_\delta(\phi^{-s} y_1)\subset \phi^l
\overline{B}_r(\phi^{-l} y_1)$, hence $\phi^{-l} y_1$ and $\phi^{-l}
y_2$ are in the same or in the neighboring tiles of ${\mathcal T}$ by the
choice of $r$. It is shown in the course of the proof of
\cite[Lem.\,3.4]{KS} that we can write $y_1-y_2 = \sum_{h=1}^l
\phi^h w_h$, where $w_h \in W$ for some finite set $W \subset
\phi^{-1} \Xi$ which depends only on the tiling ${\mathcal T}$ (a similar
statement, but without precise value of $l$ is proved in
\cite[Lemma\,4.5]{LS}). So
\begin{eqnarray*}
||f(\xi_1) - f(\xi_2)|| & = &|| f (\phi^{-s}
y_1) - f (\phi^{-s} y_2)|| = ||A^{-s} g (y_1) -
A^{-s} g (y_2)|| \\
& = & ||A^{-s} (g (y_1) - g (y_2))|| = ||A^{-s} g (y_1 - y_2)|| \\
& = & || A^{-s} g (\sum_{h=1}^l \phi^h w_h)|| = ||A^{-s} \sum_{h=1}^l A^h g(w_h)|| \\
& = & || \sum_{h=1}^l A^{h-s} g(w_h)|| \le \sum_{h=1}^l \gamma^{h-s}||g(w_h)|| \\
& \le & L' \gamma^{l-s}\,,
\end{eqnarray*}
for some $L'> 0$ independent of $l$. Notice that
$\gamma^{l-s} = (|\lambda_{\max}|^{l -s})^{\alpha}$, where
$\alpha = \frac{\log\gamma}{\log|\lambda_{\max}|}$. Thus
\begin{eqnarray*}
||f(\xi_1) - f(\xi_2)|| &\le& L'
(|\lambda_{\max}|^{l-s})^{\alpha} \\
& < & L'\left( \frac{|\lambda_{\max}|}{r} \delta \right)^{\alpha} \ \ \ \mbox{by} \ (\ref{randdelta-inequality})\\
& = & L ||\xi_1 - \xi_2||^{\alpha} \ \ \ \mbox{where} \ L:= L'
\left( \frac{|\lambda_{\max}|}{r} \right)^{\alpha}\,,
\end{eqnarray*}
and (\ref{Holder-continuity}) is proved.
\end{proof}
\medskip
Since $\mathcal{C}_{\infty}$ is dense in ${\mathbb R}^d$, we can
extend $f$ to a map $f : {\mathbb R}^d \to H$ by continuity, and moreover,
$$
f\circ \phi = A \circ f
$$
(we denote the extended function by the same symbol $f$).
This proves part (i) of Prop.~\ref{prop-KenSol}.
\begin{lemma} \label{depend-only-on-tile-type}
Let $T$ and $T + z$ be tiles in ${\mathcal T}$. Then \begin{eqnarray}
\label{translational-on-same-tiles} f(\xi +z) = f(\xi) + g(z) \ \
\mbox{for any} \ \xi \in \mbox{\rm supp}(T).\end{eqnarray}
\end{lemma}
\begin{proof} It is enough to show that (\ref{translational-on-same-tiles})
holds for a dense subset of $\mbox{\rm supp}(T)$, namely, ${\mathcal C}_\infty \cap \mbox{\rm supp}(T)$. Suppose that $\xi =
\phi^{-k} c(S)$, where $S \in \omega^k (T)$. Note that \begin{eqnarray}
\label{in-a-supertile} S + \phi^k z \in \omega^k (T + z) \subset
{\mathcal T}. \end{eqnarray} Then
\begin{eqnarray*}
f(\xi + z) &=& f(\phi^{-k} c(S) + z) =
f(\phi^{-k}(c(S) + \phi^k z)) \\
&=& f\phi^{-k}(c(S + \phi^k z)) = A^{-k} g (c(S + \phi^k z)) \\
&=& A^{-k}( g (c(S))+ g(\phi^k z)) = A^{-k} g (c(S)) + A^{-k}g \phi^k(z) \\
&=& f(\phi^{-k} c(S)) + g(z) = f(\xi) + g(z) \,.
\end{eqnarray*}
\end{proof}
Recall that $A$ is diagonalizable over ${\mathbb C}$. For $\theta \in {\em Spec}(A)$ let $p_\theta: H\to H$ be the canonical projection onto the real $A$-invariant subspace for
$A$ corresponding to $\theta$, so that we have
$$
I_H = \sum_{\theta \in {\em Spec}(A)} p_\theta.
$$
Define
\[f_\theta = p_{\theta} \circ f, \ \ \ \theta\in {\em Spec}(A).\]
Note that \begin{eqnarray} \label{sigmaEqualtoSmallSigmas} f= \sum_{\theta \in {\em Spec}(A)} f_\theta.\end{eqnarray}
Suppose that $\lambda \in Spec(\phi)$ satisfies
$|\lambda| = \gamma$.
\begin{lemma} \label{piij-Lipschitz-constant}
For $\theta \in Spec(A)$ and $a \in {\mathbb R}^d$,
\[ \left\{ \begin{array}{ll}
f_\theta|_{a + G_{\lambda}} \ \mbox{is Lipschitz} \ &
\mbox{if $ |\theta| = |\lambda|$}\,; \\
f_\theta|_{a + G_{\lambda}} \ \mbox{is constant} \ &
\mbox{if $ |\theta| > |\lambda|$} \,.
\end{array}
\right.
\]
Moreover, the Lipschitz constant is uniform in $a\in {\mathbb R}^d$ (equal to $C$ from (\ref{Lipschitz-linear-on-controlPointSet})).
\end{lemma}
\begin{proof} Let $\xi_1, \xi_2 \in a + G_{\lambda}$ for some $a \in {\mathbb R}^d$.
For any $l \in {\mathbb Z}_+$, using the norm in $H$ analogous to that in (\ref{def-norm}), so that $\|A \circ p_\theta({\bf x}) \| = |\theta|\,\|p_\theta({\bf x})\|$, we obtain
\begin{eqnarray} \label{pi-ij-inequality-on-control-points}
||f_\theta(\xi_1) - f_\theta(\xi_2)|| &=&
||(p_{\theta} \circ f) \phi^{-l} (\phi^l \xi_1) - (p_{\theta} \circ f) \phi^{-l} (\phi^l \xi_2)|| \nonumber\\
& = & ||p_{\theta}(A^{-l} f(\phi^l \xi_1) - A^{-l} f(\phi^l \xi_2))||\nonumber \\
& = & |\theta|^{-l}|| p_{\theta} (f(\phi^l \xi_1) -
f(\phi^l \xi_2))|| \nonumber \\
& \le & |\theta|^{-l}||
f(\phi^l \xi_1) - f(\phi^l \xi_2)||.
\end{eqnarray}
Note that there exist $y_1, y_2 \in \mathcal{C}$ such that
\[||\phi^l \xi_1 - y_1|| <
\delta_1 \ \ \mbox{and} \ ||\phi^l \xi_2 - y_2|| < \delta_1 \]
for some fixed $\delta_1 >0$. Since $f$ is uniformly continuous,
\[ ||f(\phi^l \xi_1) - f(y_1)|| <
\delta_2 \ \ \mbox{and} \ ||f(\phi^l \xi_2) -
f(y_2)|| < \delta_2 \] for some fixed $\delta_2 > 0$.
By the assumption on $f$, we have $||f(y_1) - f(y_2)|| \le C ||y_1 -
y_2||$. Thus
\begin{eqnarray} \label{pi-ij-controlPoints-Lipschitz}
||f(\phi^l \xi_1) - f(\phi^l \xi_2)|| & < &
||f(y_1)- f(y_2)|| + 2 \delta_2 \nonumber \\
& \le & C ||y_1 - y_2|| + 2 \delta_2 \nonumber \\
& < & C (||\phi^l \xi_1 - \phi^l \xi_2|| + 2\delta_1) + 2
\delta_2.
\end{eqnarray}
Applying (\ref{pi-ij-controlPoints-Lipschitz}) to
(\ref{pi-ij-inequality-on-control-points}), we obtain that for any
$l \in {\mathbb Z}_{+}$
\begin{eqnarray*}
\lefteqn{||f_\theta(\xi_1) - f_\theta(\xi_2)||} \\
& < & |\theta|^{-l}
\left( C ||\phi^l \xi_1 - \phi^l \xi_2|| + 2C \delta_1 + 2 \delta_2 \right) \\
& = & \frac{C |\lambda|^l}{|\theta|^l} ||\xi_1 -
\xi_2|| + \frac{1}{|\theta|^l}(2C \delta_1 + 2\delta_2)\,.
\end{eqnarray*}
Thus if $|\theta| = |\lambda|$, we have that $f_\theta|_{a+ G_{\lambda}}$ is
Lipschitz with a uniform Lipschitz constant $C$, and if $|\theta|
> |\lambda|$, we have that $f_\theta|_{a+ G_{\lambda}}$ is
constant.
\end{proof}
\begin{remark} \label{subspace-generated-by-eigensp}
{\em First note that $|\lambda| = \gamma \le {\rm min}\{|\theta| :
\theta \in Spec(A)\}$ by (\ref{eq-newk1}). The last lemma implies that for any $\xi \in {\mathbb R}^d$ and $w \in
G_{\lambda}$, the vector $f(\xi + w) - f(\xi)$ is in the subspace generated by
eigenspaces of $A$ corresponding to eigenvalues $\theta$ for
which $|\theta| = |\lambda|$. We make use of this observation to
show (\ref{phi-phi}) in Lemma~\ref{pi-i-on-E-lamimin} below.}
\end{remark}
From Lemma~\ref{piij-Lipschitz-constant} and
(\ref{sigmaEqualtoSmallSigmas}), we get the following corollary.
\begin{cor} \label{pi-i-Lipschitz}
$f|_{a + G_{\lambda}}$ is Lipschitz for any $a \in {\mathbb R}^d$.
\end{cor}
We now prove furthermore that $f$ is affine linear on $
G_{\lambda}$ slices of ${\mathbb R}^d$.
\begin{lemma} \label{pi-i-on-E-lamimin}
$f|_{a + G_{\lambda}}$ is affine linear for any $a \in {\mathbb R}^d$.
\end{lemma}
\begin{proof} This is analogous to \cite[Lem.\,3.7]{KS}, but in some places the presentation is sketchy, so we provide complete details for the
readers' convenience.
Since $f|_{a+ G_{\lambda}}$ is Lipschitz for any $a \in {\mathbb R}^d$,
it is a.e.\ differentiable by Rademacher's theorem, and hence $f$ is differentiable in the direction of $G_{\lambda}$ a.e. in $
{\mathbb R}^d$, by Fubini's theorem. Let \[ D(z)u = \lim_{t \to 0}
\frac{f(z + tu) - f(z)}{t} \ \ \ \mbox{for} \ u \in G_{\lambda} \
\mbox{and} \ z \in {\mathbb R}^d .\] The limit exists a.e.\ $z \in {\mathbb R}^d$ and
for all $u \in G_{\lambda}$, and $D(z)u$ is a linear transformation
in $u$ (from $G_{\lambda}$ to $H$). Moreover $D(z)$ is a
measurable function of $z$, being a limit of continuous functions. By the definition of total derivative,
\[ \lim_{n \to \infty} F_n(z)=0\ \ \
\mbox{for a.e.} \ z \in {\mathbb R}^d,\ \ \mbox{where}\ \
\ F_{n}(z) = \sup_{\stackrel{u
\in G_{\lambda}}{0 < \|u\| < \frac{1}{n}}} \frac{\|f(z + u) - f(z)
- D(z)u \|}{\|u\|}\,.\]
By Egorov's theorem, $\{F_n\}$
converges uniformly on a set of positive measure. This implies that
there exists a sequence of positive integers $N_l \uparrow
\infty$ such that
\begin{eqnarray} \label{define-Omega}
\Omega := \left\{\xi \in {\mathbb R}^d :\ \forall\,l\ge 1,\ \forall\, u \in B_{1/N_l}\cap G_l,\
\frac{\|f(\xi + u) - f(\xi) - D(\xi)u\|}{\|u\|} < \frac{1}{l} \right\}
\end{eqnarray} has positive Lebesgue measure.
Our goal is proving that $\Omega$ has full Lebesgue measure. The
argument is based on a kind of ``ergodicity''. First observe from
Lemma\,\ref{depend-only-on-tile-type} that $\Omega$ is ``piecewise
translation-invariant'' in the following sense: \begin{eqnarray}
\label{eq-trans} (T\in {\mathcal T},\ T+x\in {\mathcal T},\ \xi\in \Omega\cap
\mbox{\rm supp}(T))\ \Longrightarrow\ \xi + x \in \Omega. \end{eqnarray} Second,
$\Omega$ is forward invariant under the expansion map $\phi$.
Indeed, let $\xi \in \Omega$ and $u \in \phi(B_{1/N_l}) \cap
G_\lambda$. Then
\begin{eqnarray}
\lefteqn{\|f(\phi\xi +u) - f(\phi \xi) - A D(\xi) \phi^{-1} u\|} \nonumber \\ & = & \|A(f(\xi+\phi^{-1}u) - f(\xi) - D(\xi) \phi^{-1}u)\| \nonumber \\
& = & |\lambda|\|f(\xi + \phi^{-1}u) - f(\xi) - D(\xi) \phi^{-1}u)\| \ \ \
\ \ \ \ \mbox{by Remark~\ref{subspace-generated-by-eigensp}} \nonumber \\
& < & |\lambda| \frac{\|\phi^{-1}u\|}{l} = \frac{|\lambda|}{|\lambda|} \frac{\|u\|}{l} =
\frac{\|u\|}{l}\,. \label{phi-phi}
\end{eqnarray}
This implies that $D(\phi \xi)$ exists and equals $AD(\xi) \phi^{-1}$, and since $\phi(B_{1/N_l}) \supset B_{|\lambda|/N_l} \supset B_{1/N_l}$, we also obtain that
$$
\phi(\Omega) \subset \Omega.
$$
We will need a version of the Lebesgue-Vitali density theorem where the differentiation basis is the collection of sets of the form $\phi^{-l}B_1,\ l\ge 0$, and
their translates. It is well-known that such sets form a density basis, see \cite[pp.\,8--13]{Stein}. Let $y$ be a density point of $\Omega$ with respect to this density
basis. Then
\[m(\Omega \cap \phi^{-l}B_1 (\phi^l y)) \ge (1 - {\mbox{$\epsilon$}}_l) m(\phi^{-l}B_1) \ \ \ \mbox{for some} \ {\mbox{$\epsilon$}}_l \to 0, \]
where $m$ denotes the Lebesgue measure.
Note that \begin{eqnarray} m(\Omega \cap B_1 (\phi^l y)) & \ge & m(\phi^l
\Omega \cap B_1 (\phi^l
y)) \nonumber \\
& = & |\det \phi|^l m(\Omega \cap \phi^{-l} B_1 (\phi^l y)) \nonumber \\
& \ge & |\det \phi|^l (1 - {\mbox{$\epsilon$}}_l) m(\phi^{-l}B_1)\nonumber \\
& = & (1 - {\mbox{$\epsilon$}}_l) m(B_1). \nonumber \end{eqnarray} By FLC and repetitivity, there exists $R>0$ such that $B_R$ contains equivalence classes of all the patches $[B_1(\phi^l y)]^{\mathcal T}$.
Then for any $l \in {\mathbb Z}_+$, there exists $y_l \in B_R$ such that
$$
[B_1(y_l)]^{{\mathcal T}} = [B_1(\phi^l y)]^{\mathcal T} + (y_l - \phi^l y).
$$
By (\ref{eq-trans}), we have
$$
m(\Omega \cap B_1(y_l)) \ge (1 - {\mbox{$\epsilon$}}_l) m(B_1),
$$
hence $m(\Omega \cap B_1(y')) = m(B_1)$ for any limit point $y'$ of the sequence $\{y_l\}$. We have shown that $\Omega$ is a set of full measure in $B_1(y')$.
But then it is also a set of full measure in $\phi^k B_1(y')$ for $k \ge 1$. By the repetivity of ${\mathcal T}$, using (\ref{eq-trans}), we obtain that
$\Omega$ has full measure in ${\mathbb R}^d$.
Choose $n_l \in {\mathbb Z}_+$ so that
$|\lambda|^{n_l} > N_l$. Repeating the argument of (\ref{phi-phi}) we obtain
\begin{eqnarray} \xi \in \phi^{n_l} \Omega &\Rightarrow & ||f(\xi + v) - f(\xi) - D(\xi) v|| \le \frac{||v||}{l} \nonumber \\
&& \ \ \ \mbox{for all} \ v \in \phi^{n_l} \left( B_{1/{N_l}} \cap
G_{\lambda} \right) \supset B_{|\lambda|^{n_l}/N_l} \cap G_\lambda \supset
B_1 \cap G_{\lambda} \, . \nonumber \end{eqnarray} Thus $f(\xi + v) = f(\xi) +
D(\xi) v$ for any $\xi \in \bigcap_{l=1}^{\infty}
\phi^{n_l}\Omega$ and $v \in B_1 \cap G_{\lambda}$. Note that
$\bigcap_{l=1}^{\infty} \phi^{n_l} \Omega$ has full measure, hence
it is dense in ${\mathbb R}^d$. So for any $\xi \in {\mathbb R}^d$, we can find a
sequence $\{\xi_j\} \subset \bigcap_{l=1}^{\infty} \phi^{n_l}
\Omega$ such that $\xi_j \to \xi$. Since $f|_{\xi_j + G_\lambda}$ is
Lipschitz with a uniform Lipschitz constant $C$, the derivatives
$D(\xi_j)$ are uniformly bounded, and we can assume that
$D(\xi_j)$ converges to some linear transformation $D_\xi$ by
passing to a subsequence. Then we can let $j\to \infty$ to obtain
$$
f(\xi + v) = f(\xi) + D_\xi v\ \ \ \mbox{for all}\ v\in B_1 \cap G_\lambda.
$$
Since this holds for every point in ${\mathbb R}^d$, we obtain that $D_\xi = D_{\xi'}$ for any $\xi, \xi' \in {\mathbb R}^d$ with $\xi - \xi' \in G_{\lambda}$, and
$f|_{\xi+G_{\lambda}}$ is affine linear for any $\xi \in {\mathbb R}^d$.
\end{proof}
This concludes the proof of Proposition~\ref{prop-KenSol}.
\medskip
Recall (\ref{def-Geps}) that $G' = \bigcap_{{\mbox{$\epsilon$}}>0} G_{\mbox{$\epsilon$}}$ and $G_{\mbox{$\epsilon$}} = {\rm Span}_{\mathbb R}(B_{\mbox{$\epsilon$}} \cap (I-P_\lambda)\Xi)$.
\begin{lemma} \label{lem-G'}
There exists ${\mbox{$\epsilon$}}>0$ such that $G' = G_{{\mbox{$\epsilon$}}'}$ for every $0 < {\mbox{$\epsilon$}}'\le {\mbox{$\epsilon$}}$, and moreover,
$$
G' = G'':={\rm Span}_{\mathbb R}(B_{{\mbox{$\epsilon$}}'} \cap (I-P_\lambda)({\mathcal C}_1- {\mathcal C}_1))\ \ \mbox{for all}\ 0 < {\mbox{$\epsilon$}}'\le {\mbox{$\epsilon$}}.
$$
\end{lemma}
\begin{proof}
Observe that $G_{{\mbox{$\epsilon$}}'} \subset G_{\mbox{$\epsilon$}}$ for ${\mbox{$\epsilon$}}' < {\mbox{$\epsilon$}}$. These are finite-dimensional subspaces over ${\mathbb R}$, hence they must stabilize, which yields the first claim.
To prove the second claim, we just need to show $G' \subset G''$ since ${\mathcal C}_1-{\mathcal C}_1 \subset \Xi$.
There exists $k \in {\mathbb Z}_+$ such that $\phi^k\Xi \subset {\mathcal C}_1-{\mathcal C}_1$ (just choose $k$ such that
$\omega^k(T_1)$ contains tiles of all types). Then
$$
G' = \phi^k G' = \phi^k G_{{\mbox{$\epsilon$}}'/\|\phi\|^k} \subset {\rm Span}_{\mathbb R} (B_{{\mbox{$\epsilon$}}'} \cap (I-P_\lambda) \phi^k \Xi) \subset G'',
$$
as desired.
\end{proof}
\medskip
\noindent {\em Proof of Proposition~\ref{prop-KenSol2}.} This is similar to \cite[Lem.\,3.8]{KS}, but again, there are some differences, and we provide more details here.
For any $z \in {\mathbb R}^d$ and $\xi \in G$, we have $f(z+ P_{\lambda}\xi) = f(z)
+ E(z)P_{\lambda} \xi$ by Lemma \ref{pi-i-on-E-lamimin}. Since $f$ is
uniformly continuous, $E(z)$ is independent of the choice of $z$.
So for $\xi \in G$,
\begin{eqnarray} f(a + \xi) &=&
f(a + (I - P_{\lambda})\xi + P_{\lambda} \xi) \nonumber \\
&=& f(a + (I - P_{\lambda})\xi) + E \, P_{\lambda} \xi \label{pumba}
\,,
\end{eqnarray} for some fixed linear transformation $E:\,G_\lambda\to H$. Thus we only need to prove that
$f$ is affine linear on all $G'$-slices.
Let $T$ be a tile of type 1 in ${\mathcal T}$. For any $a \in
(\mbox{\rm supp}(T))^{\circ}$, choose $r
> 0$ such that $B_r(a) \subset (\mbox{\rm supp}(T))^{\circ}$. We will show that
\begin{eqnarray} \label{Jensen-functional-equation} f\left(\frac{\zeta_1 +
\zeta_2}{2}\right) = \frac{f(\zeta_1) + f(\zeta_2)}{2} \ \ \
\mbox{for all} \ \zeta_1, \zeta_2 \in B_r(a) \cap (a + G')\,. \end{eqnarray}
In other words, $f|_{a+G'}$ satisfies the so-called Jensen
functional equation, and since $f$ is continuous, this will imply
that $f|_{a+G'}$ is locally affine linear, see
\cite[2.1.4]{Aczel}. By expanding (using $\phi$-invariance) and
translating (using (\ref{eq-trans})), we will then conclude that
$f|_{a+G'}$ is affine linear for all $a \in {\mathbb R}^d$.
Now we show (\ref{Jensen-functional-equation}). By Lemma~\ref{lem-G'}, for any
${\mbox{$\epsilon$}}' > 0$ with ${\mbox{$\epsilon$}}' \le {\mbox{$\epsilon$}}$, there exists a basis $\{y_1,
\cdots, y_s \}$ of $G'$ such that for each $1 \le j \le s$, $y_j
\in B_{{\mbox{$\epsilon$}}'}$ and $y_j = (I - P_{\lambda}) z_j$ for some $z_j \in
\mathcal{C}_1 - \mathcal{C}_1$. Let
$\zeta_1, \zeta_2 \in B_r(a) \cap (a + G')$ and fix small ${\mbox{$\epsilon$}}'
> 0$ such that ${\mbox{$\epsilon$}}' \le {\mbox{$\epsilon$}}$ and
\[ {\mbox{$\epsilon$}}' < \frac{(r-\mbox{max}\{||\zeta_1||, ||\zeta_2||\})}{4s}.\] We consider the lattice
generated by the $y_j$'s in $G'$. It defines a grid with
grid cells of diameter less than $s \max_j||y_j|| \le s {\mbox{$\epsilon$}}'$.
Thus there exist $b_j \in {\mathbb Z}$, $1 \le j \le s$, such that
\[||\sum_{j=1}^s b_j y_j - \frac{(\zeta_2 - \zeta_1)}{2}|| < s
{\mbox{$\epsilon$}}'.\] Let $\tilde{\zeta} : = \zeta_1 + \sum_{j=1}^s b_j y_j$, so that $$\|\frac{\zeta_1+\zeta_2}{2}-\tilde{\zeta}\|< s{\mbox{$\epsilon$}}'.$$
Translate our grid in such a way that $\zeta_1$ is the
origin and consider a `grid geodesic' connecting $\zeta_1$ to
$\tilde{\zeta} = \zeta_1 + \sum_{j=1}^s b_j y_j $ in the $s
{\mbox{$\epsilon$}}'$-tube around the line segment $[\zeta_1, \tilde{\zeta}]$. By
the choice of ${\mbox{$\epsilon$}}'$, this `grid geodesic' is contained in
$B_r(a) \cap (a+ G')$. It is a sequence of points $\xi_1 =
\zeta_1, \xi_2, \cdots, \xi_L = \tilde{\zeta}$, where $\xi_{l+1} -
\xi_l = y_{t(l)}$, $y_{t(l)} = (I - P_{\lambda})z_{t(l)}$, $z_{t(l)}
\in \mathcal{C}_1 - \mathcal{C}_1$, and $L = \sum_{j=1}^s |b_j|$.
For each $z_{t(l)} \in \mathcal{C}_1 - \mathcal{C}_1$, there
exists a tile $S$ of type 1 such that $S + z_{t(l)} \in {\mathcal T}$. By
Lemma~\ref{depend-only-on-tile-type} we have
$f(\eta + z_{t(l)}) = f(\eta) + g (z_{t(l)})$ for any $\eta \in
\mbox{\rm supp}(S)$. In view of (\ref{pumba}),
\begin{eqnarray} \label{linearity-on-tile} f(\eta +y_{t(l)}) =
f(\eta) + g(z_{t(l)}) - E \, P_{\lambda} (z_{t(l)}) \,. \end{eqnarray}
Since $\xi_l, \xi_{l+1}\in \mbox{\rm supp}(T)$ which is of type 1, using
Lemma~\ref{depend-only-on-tile-type} again we obtain
\begin{eqnarray*}
f(\xi_{l+1}) - f(\xi_{l}) & = & f(\xi_{l} + y_{t(l)}) - f(\xi_{l}) \\
& = & f(\eta +y_{t(l)}) - f(\eta) = g(z_{t(l)}) -
E \, P_{\lambda}(z_{t(l)}).
\end{eqnarray*}
Then \begin{eqnarray} \label{puma1} f(\tilde{\zeta}) - f(\zeta_1) =
\sum_{l=1}^L (g(z_{t(l)}) - E \, P_{\lambda}(z_{t(l)})).\end{eqnarray} Note that
$\zeta_2 - (\zeta_2 + \zeta_1 - \tilde{\zeta}) = \sum_{j=1}^s b_j
y_j$. The point $\zeta_2 + \zeta_1 - \tilde{\zeta}$ is symmetric
to $\tilde{\zeta}$ with respect to $\frac{\zeta_1+\zeta_2}{2}$, so
it is also within $s{\mbox{$\epsilon$}}'$ of $\frac{\zeta_1+\zeta_2}{2}$. The
grid geodesic which connected $\zeta_1$ to $\tilde{\zeta}$,
translated by $\zeta_2-\tilde{\zeta}$, connects $\zeta_2 + \zeta_1
- \tilde{\zeta}$ to $\zeta_2$ inside $B_r(a) \cap (a+ G')$. Thus,
we obtain, repeating the argument above, that \begin{eqnarray} \label{puma2}
f(\zeta_2) - f(\zeta_2 + \zeta_1 - \tilde{\zeta}) = \sum_{l=1}^L
(g(z_{t(l)}) - E \, P_{\lambda}(z_{t(l)})).\end{eqnarray} Since $||\tilde{\zeta}
- \frac{\zeta_2 + \zeta_1}{2}|| < s {\mbox{$\epsilon$}}'$, by uniform continuity
\[\mbox{max}\{||f(\tilde{\zeta}) - f(\frac{\zeta_2 + \zeta_1}{2})||,
||f(\zeta_2 + \zeta_1 - \tilde{\zeta}) -
f(\frac{\zeta_2 + \zeta_1}{2})||\} < \delta({\mbox{$\epsilon$}}')\]
where $\delta({\mbox{$\epsilon$}}')\to 0$ as ${\mbox{$\epsilon$}}'\to 0$. Combining this with (\ref{puma1}) and (\ref{puma2}) yields
(\ref{Jensen-functional-equation}), as desired.
\hfill $\square$
\medskip
This completes the proof of Theorem~\ref{isomorphicImageOfC-contained}.
|
\section{Generalities}\label{QP}
In this section, we recall the basics of classical black holes,
introduce some notation, outline the renormalisation group improvement
for black hole metrics and discuss first implications.
\subsection{Schwarzschild metric}
The classical, static, spherically symmetric, non-charged black hole
solution to Einstein's equation is the well-known Schwarzschild black
hole \cite{Schwarzschild:1916uq}. Its line element in $d\ge 4$ dimensions is
given by \cite{Tangherlini:1963bw} (see also \cite{Myers:1986un})
\begin{equation}\label{ds2}
ds^2=-f(r)\ dt^2 +\frac{dr^2}{f(r)} + r^2\ d\Omega^2_{d-2}\,.
\end{equation}
The lapse function
\begin{equation}\label{f}
f(r)=
1-\frac{G_N\,M}{r^{d-3}}
\end{equation}
depends on Newton's
coupling constant $G_N$ in $d$ dimensions. The reduced black hole mass $M$ is
\begin{eqnarray} \label{M-def}
M=\frac{8\,\Gamma(\frac{d-1}{2})}{(d-2)\pi^{(d-3)/2}}\, M_{\rm phys}\,,
\end{eqnarray}
with $M_{\rm phys}$ the physical mass of the black hole.
In terms of these, the classical Schwarzschild radius $r_{\rm cl}$ is given as
\begin{eqnarray} \label{classicalBH-def}
r_{\rm cl}^{d-3}&=& G_N \, M\,.
\end{eqnarray}
The black hole solution is continuous in the mass parameter $M$ and
displays a Bekenstein-Hawking temperature inversely proportional to
its mass. For large radial distance $r\to\infty$, we observe $f(r)\to
1$, indicating that the geometry of a Schwarzschild space-time becomes
flat Minkowskian. The coordinate singularity at $r=r_{\rm cl}$ where
$f(r_{\rm cl})$ vanishes, defines the event horizon of the black hole.
In the short distance limit $r\to 0$ we observe a divergence in
$f(r)$, reflecting a metric and curvature singularity at the origin.
\begin{center}
\begin{table*}[t]
\begin{tabular}{c|c|c|c|c}
case
& short distance index
& gravity
& horizons
&${}$\quad $f(r\to 0)$${}$\quad${}$ \\
\hline
(i)
& ${}$\quad $\alpha<d-3$ \quad ${}$
& ${}$\quad strong, if $\alpha<0$;\quad weak, if $\alpha>0$${}$\quad ${}$
& one
& ${}$\quad singular${}$\quad\\[.5ex]
(ii)
&${}$\quad $\alpha=d-3$\quad ${}$
& weak
& none, one or more
& finite\\[.5ex]
${}$\quad${}$ (iii)\quad${}$
&$\alpha>d-3$
& weak
& none, one or more
&1\\
\end{tabular}
\caption{\label{T1} Horizons of quantum-corrected Schwarzschild black holes assuming a
scale-dependent gravitational coupling strength \eq{G(r)1} at short distances for various
dimensions and in dependence on the short distance index $\alpha$ (see text).}
\end{table*}
\end{center}
\subsection{Improved metric}\label{IM}
The classical black hole is modified once quantum gravitational
effects are taken into account. In general, quantum fluctuations will
modify the gravitational force law by turning Newton's coupling $G_N$
into a distance-dependent ``running" coupling $G(r)$,
\begin{equation}
G_N\to G(r)\,.\label{RGimproved}
\end{equation}
It is the central assumption of this paper that the leading quantum
gravitational corrections to the black hole are captured by the
replacement \eq{RGimproved} in the metric \eq{f}. This
``renormalisation group improvement" should provide a good description
of the leading quantum corrections, because the primary, explicit,
dependence of the Schwarzschild black hole on the gravitational sector
is only via Newton's coupling $G_N$. Furthermore, the classical black
hole solution is continuous in its mass parameter $M$, and the effects
of quantum corrections are parametrically suppressed for large black
hole mass with $M_D/M$ serving as an external, small, control
parameter.
Whether gravity becomes ``strong" at shortest distances, or ``weak",
will depend on the ultraviolet completion for gravity and the related
running under the renormalisation group.
Next we discuss the main
implications arising from a running gravitational
coupling. For the sake of the argument, we parametrize $G(r)$ as
\begin{equation}\label{G(r)1}
G(r)=r_{\rm char}^{d-2} \left(\frac{r}{r_{\rm char}}\right)^{\alpha}
\end{equation}
for sufficiently small $r$, where $r_{\rm char}$ denotes a characteristic
length scale where quantum corrections become dominant. The index
$\alpha$ then parametrizes the gravitational coupling strength at
short distances, with $\alpha>0$ $(\alpha<0)$ denoting a decrease
(increase) of $G(r)/G_N$ at small distances, respectively, and the
classical limit $\alpha=0$ where $r_{\rm char}$ is given by the Planck
length $r_{\rm char}=1/M_D$. The behaviour of $f(r\to 0)$, and the
solutions to the horizon condition $f(r)=0$ then teach us how the
RG-improved black hole depends on the quantum effects parametrized by
$\alpha$. The qualitative pattern is summarised in Tab.~\ref{T1}. We distinguish
three cases, depending on the short distance index $\alpha$:
\begin{itemize}
\item[(i)] $\alpha<d-3$. In this case the gravitational coupling
either increases with decreasing $r$, or even decreases slightly,
though not strongly enough to overcome the enhancement due to the
$\s01{r^{d-3}}$-factor in \eq{f}. Therefore $f(r)$ unavoidably has
to change sign leading to a horizon. This includes the classical
case $\alpha=0$, and all cases of strong gravity corresponding to a
diverging $G(r)/G_N$ for small $r$. Interestingly, even if gravity
weakens at short distances with an index $0<\alpha<d-3$, we still
observe a horizon for arbitrary small black hole masses.
\item[(ii)] $\alpha=d-3$. In this case, we have a finite limit $f(r\to
0)=f_0$. For $f_0<0$, this necessarily enforces a horizon, similar
to case (i). For $f_0>0$, the situation is analogous to case (iii).
\item[(iii)] $\alpha>d-3$. In this case, $G(r)$ weakens fast enough to
overcome the enhancement due to $\s01{r^{d-3}}$. Therefore $f(r\to
0)\to 1$ and $f(r)$ may display either several zeros, a single one,
or none at all, leading to several, one or no horizon depending on
the black hole mass $M$ and the precise short-distance behaviour of
$G(r)$.
\end{itemize}
We conclude that for $\alpha >d-3$ the Schwarzschild black hole may no
longer display a horizon for all mass, whereas for $\alpha<d-3$ a
horizon is guaranteed for all $M$.
Which of these scenarios is realised depends on the short-distance
behaviour of gravity.
In the remaining part of the paper we access this picture
quantitatively, using the renormalisation group for gravity.
\section{Asymptotically safe gravity} \label{QG}
In this section, we discuss field theory based approaches to quantum
gravity including effective theory and the asymptotic safety scenario for gravity,
and provide the renormalisation group running for Newton's coupling.
\subsection{Effective theory for gravity}\label{ET}
In the absence of a complete theory for quantum gravity, quantum corrections
of the form \eq{RGimproved} can be accessed in the weak gravity regime
using methods from effective theory \cite{Donoghue:1993eb,Burgess:2003jk}.
In practice, this amounts to an ultraviolet regularisation of the theory by an
UV cutoff $\Lambda$ of the order
of the fundamental Planck scale. In the weak gravity regime where
$r\, M_D\gg 1$ with Planck mass $M_D=(G_N)^{-1/(d-2)}$, it has been
found that
\begin{equation}
{G(r)}={G_N}\left(1-\frac{\omega\,G_N}{r^2}\right)
\end{equation}
in four dimensions, and at the one-loop order
\cite{Hamber:1995cq,Bjerrum-Bohr:2002ks,Bjerrum-Bohr:2002kt,Akhundov:2006gh},
with $\omega>0$ (see \cite{Duff:1974ud} for earlier results). In higher dimensions, no effective theory
results are available and
thus we have to provide the relevant RG input from a different source.
\subsection{Asymptotic safety}\label{AS}
Asymptotic safety of gravity is a scenario where gravity exists as a well-defined
fundamental local quantum theory of the metric field \cite{Weinberg}. This set-up goes one step
beyond an effective theory approach: it assumes that the
ultraviolet cutoff $\Lambda$ from effective theory can safely be removed,
$\Lambda\to\infty$, whereby the relevant gravitational couplings approach
a non-trivial fixed point \cite{Litim:2008tt,Litim:2006dx,Niedermaier:2006ns,Niedermaier:2006wt,Percacci:2007sz}.
We briefly recall the main picture. Consider
the dimensionless gravitational coupling
\begin{equation}
g(\mu)=G(\mu)\mu^{d-2}\equiv G_0 Z_G^{-1}(\mu)\mu^{d-2}\,,
\end{equation}
where $\mu$ denotes the RG momentum scale,
$G(\mu)$ is the running Newton coupling,
and $Z_G$ the gravitational wave function factor.
The wave function factor is normalised as
$Z_G(\mu_0)=1$ at some reference scale $\mu_0$ with $G(\mu_0)$ given
by Newton's constant $G_0\equiv G_N$.
Then the gravitational Callan-Symanzik equation reads \cite{Litim:2003vp,Niedermaier:2006ns}
\begin{equation}\label{dg}
\frac{{\rm d}g(\mu)}{{\rm d}\ln \mu} = (d-2+\eta)g(\mu)\,.
\end{equation}
Here $\eta=-\mu\partial_\mu \ln Z_G$ denotes the anomalous dimension
of the graviton. In general, the anomalous
dimension depends on all couplings of the theory. Due to its
structure, \eq{dg} can achieve two types of fixed points. At small
coupling, the anomalous dimension vanishes and $g=0$ corresponds to
the non-interacting ($i.e.$~Gaussian) fixed point of \eq{dg}. This
fixed point dominates the deep infrared region of gravity
$\mu\to 0$. In turn, an interacting fixed point $g_*$ is achieved
if the anomalous dimension of the graviton becomes non-perturbatively
large,
\begin{equation}\label{eta}
\eta_* = 2-d\,.
\end{equation}
A non-trivial fixed point of quantum gravity in $d>2$ implies a
negative integer value for the graviton anomalous dimension,
counter-balancing the canonical dimension of $G$. As a consequence,
$G(\mu)\to
g_*/\mu^{d-2}$ in the vicinity of a non-trivial fixed point. In the UV
limit where $\mu\to \infty$, the gravitational coupling $G(\mu)$
becomes arbitrarily weak.
\subsection{Renormalisation group}\label{RG}
The above picture is substantiated through explicit renormalisation group studies for
gravity. A powerful tool is given by the functional renormalisation group \cite{ERG-Reviews},
based on the idea of integrating-out momentum modes from a path-integral
representation of quantum field theory \cite{Niedermaier:2006ns,Niedermaier:2006wt,Litim:2008tt}.
The corresponding `flowing' effective action for gravity reads
\begin{equation}\label{EHk}
\Gamma_k=
\0{1}{16\pi G_k}\int d^dx \sqrt{g}\left[-R(g)+\cdots\right]\,.
\end{equation}
Here $R(g)$ denotes the Ricci scalar, and the dots in \eq{EHk}
stand for the cosmological constant, higher dimensional operators in
the metric field, gravity-matter interactions, a classical gauge
fixing and ghost terms. Furthermore, all couplings in \eq{EHk}
are `running' couplings as functions of the Wilsonian momentum scale
$k$, which now takes over the role of the RG scale $\mu$ discussed
above. The action \eq{EHk} reduces to the standard quantum effective action
in the limit $k\to 0$, where all quantum fluctuations are integrated out.
For $k\ll M_D$, the gravitational sector is well-approximated by
the Einstein-Hilbert action with $G_k\approx G_0$, and similarly for
the gravity-matter couplings. The corresponding operators scale
canonically. At $k\approx M_D$ and above, the non-trivial RG running
of gravitational couplings becomes important.
An exact functional flow equation which governs the $k$-dependence for an action \eq{EHk}
has been put forward by Wetterich \cite{Wetterich:1992yh},
\begin{equation}\label{ERG}
\partial_t \Gamma_k=
\frac{1}{2} {\rm Tr} \left({\Gamma_k^{(2)}+R_k}\right)^{-1}\partial_t R_k
\end{equation}
and $t=\ln k$. The trace stands for a momentum integration and a sum
over indices and fields, and $R_k(q^2)$ denotes an appropriate
infrared cutoff function at momentum scale $q^2\approx k^2$
\cite{Litim:2000ci}. The flow \eq{ERG} can be seen as a functional Callan-Symanzik equation, where the mass term is replaced by a momentum-depedent mass $k^2\to R_k(q^2)$ \cite{Symanzik:1970rt}. When implemented for gravity \cite{Reuter:1996cp},
an additional background field is introduced to achieve diffeomorphism
invariance within the background field technique \cite{ERG-Reviews,Freire:2000bq,Reuter:1996cp,Litim:2002ce}.
By now, a large number of papers have shown the existence
of a non-trivial fixed point for gravity including renormalisation group studies in four and higher dimensions
\cite{Reuter:1996cp,Souma:1999at,Lauscher:2001ya,Lauscher:2002mb,Reuter:2001ag,Percacci:2002ie,Litim:2003vp,Bonanno:2004sy,Percacci:2005wu,Fischer:2006fz,Fischer:2006at,Codello:2006in,Codello:2007bd,Machado:2007ea,Rahmede:2009,Benedetti:2009gn,Niedermaier:2009zz,Percacci:2009dt},
and numerical simulations on the lattice
\cite{Hamber:1999nu,Ambjorn:2004qm}.
Analytical results for the running of the gravitational coupling have been given in \cite{Litim:2003vp}, where
\eq{EHk} has been approximated by the Ricci scalar. The central result is not altered through
the inclusion of a cosmological constant \cite{Fischer:2006fz}. Using \eq{EHk} and \eq{ERG},
one finds
\begin{eqnarray}\label{betag0}
\beta_g&=&\0{(1-4dg/c_d)(d-2)g}{1-(2d-4)g/c_d}
\end{eqnarray}
with parameter
$c_d=(4\pi)^{d/2-1}\Gamma(\s0d2+2)$. The scale-dependence of the anomalous dimension
is given via the scale-dependence of the running coupling,
\begin{eqnarray}
\eta&=&\0{2(d-2)(d+2)\,g/c_d}{2(d-2)\,g/c_d-1}\,.
\end{eqnarray}
We observe a Gaussian fixed
point at $g_*=0$ and a non-Gaussian one at $g_*=c_d/(4d)$.
Integrating the flow \eq{betag0} gives an implicit equation for $G_k$,
\begin{equation}\label{g0-explicit}
\0{G(k)}{G(k_0)}
=\left(\0{g_*-G(k)\,k^{d-2}}{g_*-\,G(k_0)\,k_0^{d-2}}\right)^{(d-2)/\theta}
\end{equation}
with boundary condition $G(k_0) k_0^{d-2}< g_*$, and the non-perturbative scaling exponent $\theta=2d\
\0{d-2}{d+2}$. The fixed point value and the scaling exponent depend slightly on the underlying momentum cutoff \cite{Litim:2003vp,Fischer:2006fz}. Inserting the running coupling
\eq{g0-explicit} into \eq{betag0} shows that the anomalous dimension displays
a smooth cross-over between the IR domain $k\ll M_D$ where $\eta \approx 0$
and the UV domain $k\gg M_D$ where $\eta \approx2-d$.
The cross-over regime becomes narrower with increasing
dimension. For our purposes, it will be sufficient to approximate the non-perturbative solution \eq{g0-explicit} further by setting the scaling index $\theta$ to $\theta=d-2$. In the limit where $G(k_0)k^{d-2}_0\ll 1$, we find
\begin{equation}\label{G(k)}
\frac{1}{G(k)}=\frac{1}{G_0} +\omega \,k^{d-2}
\end{equation}
where $ \omega=1/g_*$ is a positive constant, and $G_0=G(k_0=0)$. Note that \eq{G(k)} looks, formally, like a 1-loop equation. The difference here is that the coefficient $\omega$, in general, also encodes information about the underlying fixed point and may be numerically different from the 1-loop coefficient. This equation captures the main
cross-over behaviour.
\begin{figure*}
\subfigure[\label{disdim}\ Proper and linear distance in various dimensions.]{
\includegraphics[width=.45\hsize]{disdimOct.eps}}
\subfigure[\label{LogDis}\ Several distance functions in $d=7$ dimensions.]{
\includegraphics[width=.45\hsize]{LogLogDistances7dOct.eps}}
\caption{\label{LogDis-compare} Comparison of various distance functions $D(r)$ as functions of $r/r_{\rm cl}$. (a) Proper distance in $d=4,6,7$ and $10$ dimensions (top to bottom) and linear matching (straight line). (b) Interpolating expressions \eq{D-interpol} and \eq{D-interpol2}, proper distance matching \eq{distance}, and linear matching \eq{linear} (bottom to top) in 7 dimension. }
\end{figure*}
\subsection{Relevant scales}\label{Scales}
In order to implement quantum corrections to the classical Schwarzschild
black hole geometry, we replace the classical coupling $G$ by an $r$-dependent running coupling $G(r)$ under the RG flow. The renormalisation group provides us with a momentum-scale dependent $G(k)$. This requires, additionally, a scale identification between the momentum scale $k$ and the coordinate variable $r$ of the form
\begin{equation}
\label{match}
k(r)=\xi/D(r)\,,
\end{equation}
such that \begin{equation} \label{G(r)2}
\frac{1}{G(r)}=\frac{1}{G_0}+\frac{\omega\,\xi^{d-2}}{D^{d-2}(r)}\,.
\end{equation} The distance function
$D(r)$ should be an appropriately chosen length scale which
may depend on other parameters such as eg.~the black hole mass $M$. In
general, the matching coefficient $\xi$ is non-universal and its
numerical value will depend on the RG scheme used to obtain the RG
running of $G(k)$. In a fixed RG scheme, and for a given choice for
$D(r)$, $\xi$ can be computed explicitly using methods discussed in
eg.~\cite{Donoghue:1993eb}.
Next we introduce a variety of distance functions motivated by the
Schwarzschild metric, flat space metric, dimensional analysis, and
interpolations.
$\bullet$ {\it Dimensional analysis.} The gravitational force on a
test particle in a Schwarzschild space-time depends on two independent
dimensionful parameters, the horizon $r_{\rm cl}$ (or the black hole
mass, respectively) and the radial distance scale $r$. Therefore,
dimensional analysis suggests that a general length scale $D(r)$ can
be written as
\begin{equation}
\label{nonlinear}
D_{\rm da}(r)= c_{\gamma} r^{\gamma}\, r_{\rm cl}^{-\gamma+1}\,.
\end{equation}
for some $\gamma$, and $c_{\gamma}$ a positive constant. Moreover,
$\gamma$ may depend on dimensionless ratios such as $r/r_{\rm cl}$. An
ansatz taking into account the flat-space limit for $r\to \infty$, and
the deep Schwarzschild regime $r\ll r_{\rm cl}$, is given by
\begin{equation}
\label{da}
D_{\rm da}(r)\propto \left\{
\begin{array}{cc} r & {\rm for}\ \ r>r_{\rm cl}\\
r^{\gamma}\, r_{\rm cl}^{-\gamma+1} & {\rm for}\ \ r<r_{\rm cl}
\end{array}
\right.
\end{equation}
with coefficient $\gamma$. In the parametrisation \eq{G(r)1}, this
corresponds to the short-distance index $\alpha=\gamma(d-2)$. For
$\gamma>1$ $(\gamma<1)$, the matching enhances (counteracts) the RG
running of \eq{G(k)}. We note that $1/\gamma\to 0$ corresponds to a
decoupling limit where gravity is switched-off at scales below $r_{\rm cl}$.
$\bullet$ {\it Proper distance.} A different matching is obtained by
identifying the RG momentum scale $k$ with the inverse diffeomorphism
invariant distance $D_{\rm diff}(r)^{-1}$ \cite{Bonanno:2000ep}. Such
a distance is defined through the line integral
\begin{equation}\label{Ddef}
D_{\rm diff}(r)=\int_C \sqrt{|ds^2|} \,,
\end{equation}
where $C$ is an appropriately chosen curve in spacetime. Using the
classical Schwarzschild metric, we consider a path $C$ running
radially from $0$ to $r$, thereby connecting time-like with space-like
regions. With $dt=d\Omega=0$ this defines the proper distance
\begin{equation}\label{distance}
D_{\rm Schw}(r)=\int_{0}^{r} dr
\left|1-\left(\frac{r_{\rm cl}}{r}\right)^{d-3}\right|^{-1/2}\,.
\end{equation}
For any $d$, \eq{distance} has an integrable pole $\sim
1/\sqrt{r-r_{\rm cl}}$ at the connection point between space-like and
time-like regions $r=r_{\rm cl}$. Analytical expressions for $D_{\rm
Schw}(r)$ are obtained from \eq{distance} for fixed dimension. We
note that \eq{distance} corresponds to \eq{nonlinear} with an
$(r/r_{\rm cl})$-dependent index $\gamma$. For large $r\gg r_{\rm
cl}$, we have $D_{\rm Schw}(r) \rightarrow r$, where the
Schwarzschild metric becomes flat corresponding to $\gamma=1$ in
\eq{nonlinear}. For small $r$ \eq{distance} corresponds to
\eq{nonlinear} with $\gamma=(d-1)/2$.
$\bullet$ {\it IR matching.}
If the black hole mass $M$ is sufficiently large compared to the
Planck mass, we can assume that the only RG relevant length scale in the problem
is given by $r$. Therefore, $r$ is directly
identified with the (inverse) RG scale $k$,
\begin{equation}
\label{linear}
D_{\rm ir}(r)=r\,.
\end{equation}
This matching \eq{linear} corresponds to \eq{nonlinear}
with $\gamma=c_{\gamma}=1$. In the parametrisation \eq{G(r)1}, the short distance behaviour corresponds to
$\alpha=d-2$. We therefore expect the matching
\eq{linear} to capture the leading quantum effects correctly.
$\bullet$ {\it UV matching.}
For small $r\ll r_{\rm cl}$, the proper distance $D_{\rm Schw}(r)$ scales like a power-law in $r$. We find
\begin{equation}\label{D-UV}
D_{\rm uv}(r)=\frac{2\,r^{(d-1)/2}}{(d-1)\,r_{\rm cl}^{(d-3)/2}}\,.
\end{equation}
Matching the RG momentum scale with the inverse proper distance \eq{D-UV} leads to
\eq{nonlinear} with $\gamma=c_\gamma^{-1}=(d-1)/2$. In the parametrisation \eq{G(r)1}, this
corresponds to the short-distance index $\alpha=(d-1)(d-2)/2>0$, which for all
$d>3$ is larger than the index $(d-2)$ obtained from linear matching. \\
$\bullet$ {\it Interpolations.}
For the
subsequent analysis, it is useful to have approximate expressions for $D_{\rm Schw}(r)$
\eq{distance} which interpolate properly between \eq{linear} and \eq{D-UV}.
We use a simple interpolation formula for general
dimension to implement the non-linear matching \eq{distance} into \eq{G(k)}
and write
\begin{eqnarray}\label{D-interpol}
D_{\rm int1}(r)&=&\frac{2\,r^{(d-1)/2}}{(d-1)\,(r_{\rm cl}+\epsilon\,r)^{(d-3)/2}}\\
\label{nonepsilon}
\epsilon&=&\left(\s0d2-\s012\right)^{-2/(d-3)}
\end{eqnarray}
which is exact for $r\to \infty$ and $r\to 0$, and $\epsilon \in [\s049,1]$
for $d \in [4,\infty]$. Alternatively, we also use
\begin{eqnarray}\label{D-interpol2}
D_{\rm int2}(r)&=&\frac{r}{1+\s012(d-1)\,(r_{\rm cl}/r)^{(d-3)/2}}\,.
\end{eqnarray}
In Fig.~\ref{LogDis-compare} we compare different
distance functions. In Fig.~\ref{disdim}, the
functions \eq{distance} are compared with the linear matching
\eq{linear} in various dimensions. For large $r$ the proper distance
\eq{distance} approaches $r$ for all $d \geq 4$. As $r/r_{\rm cl}
\searrow1$ we observe that the gradient steepens rapidly due to the
presence of an integrable pole $~ 1/ \sqrt{r-r_{\rm cl}}$. For $r \ll
r_{\rm cl}$ the gradients of each curve are steeper with increasing
$d$ due to an additional dimensional suppression in \eq{distance}. In
Fig.~\ref{LogDis} we fix $d=7$ and observe that for large $r$ the
matchings \eq{distance}. \eq{D-interpol} and \eq{D-interpol2} approach
the correct IR behaviour \eq{linear}. For small $r$ these matchings
approach the UV matching \eq{D-UV}. We also observe that the rapid
steepening of \eq{distance} around $r/r_{\rm cl}$ implies that the
transition between IR and UV behaviour is well approximated by \eq{da}.
Finally, we provide a link with the discussion of
Tab.~\ref{T1}, see Sec.~\ref{IM}. For the distance functions motivated by the
Schwarzschild metric \eq{distance}, \eq{D-UV}, \eq{D-interpol} and
\eq{D-interpol2}, we find the index $\alpha= \frac{1}{2}
(d-1)(d-2)>(d-3)$ for all $d \geq 4$, corresponding to case (iii). In
the same vein, for $D(r)$ motivated by the flat space metric
\eq{linear} we find $\alpha=d-2$ equally corresponding to case (iii).
Finally, the distance function motivated by dimensional analysis \eq{da} contains a free
parameter $\gamma$, whose natural value is of order one. It leads to
the index $\alpha=\gamma(d-2)$ and hence relates to case (iii) for all
$\gamma>\gamma_c$, where
\begin{equation}\label{gammac}
\gamma_c=\frac{d-3}{d-2}\,.
\end{equation}
We conclude that for all physically motivated distance functions
we are lead to
the scenario described by case (iii) in Tab.~\ref{T1}, independently
of the scale identification $k=k(r)$. This, therefore, appears to be a
robust prediction from the renormalisation group running implied
within asymptotically safe gravity.
\begin{figure*}
\subfigure[\label{6dlin} \ RG running with \eq{f-L} and $\gamma=1$ in $6$ dimensions.]
{\includegraphics[width=.45\hsize]{6dPlot2.eps}}
\hskip0.09\hsize
\subfigure[\label{nonlin7} \ RG running with \eq{f-non} and $\gamma=3$ in $7$ dimensions.]
{\includegraphics[width=.45\hsize]{newnonlin7Oct.eps}}
\caption{\label{RG-f-compare} Mass and renormalisation group dependence of the
RG improved function $f(x)$ with $x=r/r_{\rm cl}$ in higher dimensions.
From top to bottom: absence of horizon $\Omega > \Omega_c$, critical black hole $\Omega =
\Omega_c$, semi-classical black hole $\Omega < \Omega_c$, and classical black hole $\Omega=0$. }
\end{figure*}
\section{Asymptotically safe black holes}\label{QBH}
In this section, we implement the renormalisation group improvement and analyse the
resulting black holes, their horizon structure, and critical
Planck-size mini-black holes. The asymptotically safe black hole is obtained
by replacing $G_N$ with the running $G(r)$ \eq{G(r)2} in
\eq{classicalBH-def} and \eq{f}, leading to the improved, asymptotically safe,
lapse function
\begin{equation}\label{f-RG}
f(r,M)=1-G(r,M)\, \frac{M}{r^{d-3}} \,.
\end{equation}
At this point we make two observations. The improved Schwarzschild
radius $r_s(M)$ is given by the implicit solution of
\begin{equation}
\label{rs_Implicit} r_s^{d-3}(M)=M\,G(r_s(M),M)\,.
\end{equation}
If
\eq{f-RG} has a solution $f(r_s(M),M)=0$, then it follows that the
quantum-corrected horizon is smaller than the classical one
$r_{s}(M)<r_{\rm cl}(M)$. This is a direct consequence of
${G(r,M)}/{G_N}\le 1$ for all $r$. Secondly, if ${G(r,M)}/{G_N}$
decreases too rapidly as a function of $r$, $f(r_s(M),M)=0$ will no
longer have a real solution $r_s(M)\ge 0$, implying the absence of a
horizon.
\subsection{Horizons}\label{horizons}
To see the above picture quantitatively, we analyse the horizon
condition analytically, also comparing various matching conditions.
For a general matching the dimensional analysis \eq{nonlinear} leads
to a running Newton's constant \eq{G(r)2} of the form
\begin{equation}
\frac{G_0}{G(r)}= 1+\frac{\tilde{\omega}\,G_{0}}{r_{\rm
cl}^{d-2}}\left(\frac{r_{\rm cl}}{r}\right)^{\gamma(d-2)}
\end{equation}
with $r_{\rm cl}$ as in \eq{classicalBH-def} and \begin{equation}
\tilde{\omega}=\omega(\xi/c_\gamma)^{d-2} \end{equation} This leads to a lapse
function given by
\begin{equation} \label{gen-f}
f(x)=1-\frac{1}{x^{d-3}}\frac{x^{\gamma(d-2)}}{x^{\gamma(d-2)}+\Omega}\,,
\end{equation}
where we have also introduced the variables
\begin{eqnarray}\label{sol-RG-2}
x&=&r/r_{\rm cl}\\
r_{\rm cl}&=&( M\,G_0)^{1/(d-3)}\\
\label{sol-RG-3}
\Omega&=&\tilde\omega\, (M_D/ M)^{\frac{d-2}{d-3}}\,.
\end{eqnarray}
We define the $d$-dimensional Planck mass as $G_0=M_D^{2-d}$
corresponding to the convention used by Dimopoulos and Landsberg
\cite{Dimopoulos:2001hw}. The parameter $\Omega$ captures the RG
running of Newton's coupling, and the mass and matching parameter
dependence. The classical black hole corresponds to $\Omega=0$ which
is achieved in the limit of vanishing quantum corrections $\omega\to
0$ or in the limit of infinite black hole mass $M\to\infty$.
Therefore, the horizon condition $f(x)=0$ always includes the
classical solution $x=1$ at $r=r_{\rm cl}$ for $\Omega=0$.
\begin{figure*}
\subfigure[\label{gammaOmega} \ Horizons $r_s$ as a function of
$\Omega/\Omega_c$ for $d=7$.
Upper/ lower branch are event/ Cauchy
horizon]{\includegraphics[width=.45 \hsize]{gammaOmega.eps}}
\hskip.09\hsize
\subfigure[\label{gammaf} Shape of $f(x)$ with $x=r/r_{\rm cl}$ at criticality $\Omega=\Omega_c$.
The minima denote the degenerate horizon $x_c$.
]{\includegraphics[width=.45 \hsize]{gammaf.eps}}
\caption{\label{gamma-dependence}Dependence of Cauchy and event horizon, and of the metric coefficient $f(x)$ at criticality $\Omega=\Omega_c$, on the parameter $\gamma=1,\s065,2,3,4$, and $10$ [panel (a): from bottom to top, panel (b): from left to right] in $d=7$ dimensions. Metric singularities are absent for $\gamma\ge\gamma_{\rm dS}=\frac{d-1}{d-2}$. Large values of $\gamma\to\infty$ represent the decoupling limit (see text).}
\end{figure*}
For simplicity, we begin with the case $\gamma=1$ corresponding to the
IR matching \eq{linear}, where $f(x)$ takes the form
\begin{equation}\label{f-L}
f(x)=1-\frac{x}{x^{d-2}+\Omega}\,.
\end{equation}
For $\Omega>0$, the horizon condition becomes
\begin{equation}\label{sol-RG-1}
0=x^{d-3}-1+\Omega/x\,.
\end{equation}
We find three qualitatively different solutions, depending on the
value of $\Omega$ (see Fig.~\ref{6dlin}). In general, \eq{sol-RG-1}
has $d-2$ possibly complex roots $x(\Omega)$. For sufficiently small
$\Omega$, two of these are positive real with
$0<x_-(\Omega)<x_+(\Omega)\le 1$ and correspond to a Cauchy horizon
$x_-\equiv r_w/r_{\rm cl}$ and an outer horizon $x_+\equiv r_s/r_{\rm
cl}$. In even dimensions, the remaining roots are complex congugate
pairs, whereas in odd dimensions, one of the remaining roots is real
and negative. Analytical solutions are obtained for $x_\pm(\Omega)$ as
a power series in $\Omega$ for any $d>3$. With increasing $\Omega$
(decreasing black hole mass $M$), real solutions to \eq{sol-RG-1}
cease to exist for $\Omega>\Omega_c$. Hence, black hole solutions are
restricted to masses $M$ with
\begin{equation}\label{Oc}
\Omega\le \Omega_c\quad{\rm and}\quad
M\ge M_c\,.
\end{equation}
For a black hole of critical mass $M_c$ we find $\Omega(M_c)=
\Omega_c$. For such a critical black hole the inner Cauchy horizon and
the outer event horizon coincide, $x_-=x_+=x_c$ with a radius of $r_c \equiv
x_c \, r_{\rm cl}(M_c)$. Solving $f(x_c)=0$ and $f'(x_c)=0$
simultaneously leads to the critical parameter
\begin{eqnarray}\label{acr}
\Omega_{c}&=&(d-3)\,(d-2)^{-\frac{d-2}{d-3}}\\
\label{xcr}
x_{c}&=& (d-2)^{-1/(d-3)}\,.
\end{eqnarray}
We note that \eq{acr} is of order one for all $d\ge 4$.
Next, we consider the distance function \eq{D-interpol} whose index $\gamma$
interpolates between $\gamma=1$ for large $r$ and
$\gamma=\frac{1}{2}(d-1)$ for small $r$, similar to the matchings
\eq{distance} and \eq{D-interpol2}.
We find
\begin{equation}\label{G-LO}
G(r)=
\frac{G_0\, r^\alpha}{r^\alpha+\tilde\omega\,G_0\, (r_{\rm cl}+\epsilon\,r)^{\alpha+2-d}}
\end{equation}
with $r_{\rm cl}$ and $\epsilon$ from \eq{classicalBH-def} and \eq{nonepsilon}, and
\begin{eqnarray}
\label{nonalpha2}
\alpha&=&\frac{1}{2}(d-1)(d-2)\\
\label{nonomegatilde2}
\tilde{\omega}&=&\omega\, \xi^{d-2}\,\left(\s0d2-\s012\right)^{d-2}\,.
\end{eqnarray}
Consequently,
\begin{equation}\label{f-non}
f(x)=1-\frac{x^{\alpha-d+3}}{x^{\alpha}+\Omega(1+x \epsilon)^{\alpha-d+2}}
\end{equation}
and the horizon condition becomes
\begin{equation}\label{horizoninterpol}
x^{\alpha-d+3}=x^{\alpha}+\Omega(1+x \epsilon)^{\alpha-d+2} \,.
\end{equation}
In Fig.~\ref{nonlin7} we plot \eq{f-non} in $d=7$ for three values of
$\Omega$. The main difference in comparison with Fig.~\ref{6dlin} is
that the limit $f\to 1$ is achieved more rapidly.
Finally, we come back to a matching with general index $\gamma$,
\eq{gen-f}, with horizon condition \begin{equation} \label{nonlinhorizon}
0=1-x^{3-d}+\Omega x^{-\gamma(d-2)}\,. \end{equation} Again, three types of
solution are found for $\gamma > \frac{d-3}{d-2}$, corresponding to
two horizons ($x_+$ and $x_-$) for $\Omega<\Omega_c$, none for
$\Omega>\Omega_c$ and a single horizon ($x_+=x_-=x_c$) for
$\Omega=\Omega_c$. Solving $f(x_c)=0$ and $f'(x_c)=0$ simultaneously
leads to the critical parameter
\begin{eqnarray} \label{genxc}
x_c&=&\left(1-\frac{d-3}{\gamma(d-2)}\right)^{\frac{1}{d-3}}\\
\Omega_c&=&\frac{d-3}{\gamma(d-2)}
\left(1-\frac{d-3}{\gamma(d-2)}\right)^{\frac{\gamma(d-2)}{d-3}-1}\,.
\end{eqnarray}
It follows that condition \eq{Oc} will hold independently of the
matching used.
Next we discuss the quantitative differences between the various
distance functions. This relates to the limit $r\to 0$, where $f(r)$ approaches
$f\to 1$, though with different rates, see Figs.~\ref{RG-f-compare}.
Effectively, the rate is parametrised through
$\gamma$. We recall that the limit $\gamma \rightarrow \infty$
switches off gravity below the horizon $x_c$. This entails, in
\eq{genxc}, that $x_c \rightarrow 1$. This is nicely seen in
Fig.~\ref{gammaOmega} where the horizons are plotted as a function of
$\Omega/\Omega_c$ for various $\gamma$ with fixed dimensionality
$d=7$. In Fig.~\ref{gammaf}, instead, we use \eq{da} with $\gamma=1$
for $r>r_{\rm cl}$ and $\gamma>1$ for $r<r_{cl}$. At $\Omega=\Omega_c$
we show $f(r)$ for various $\gamma$, and note that the limit $f \to 1$
is approached more rapidly for larger values of $\gamma$, as
expected. We conclude that $\gamma>1$ enhances the weakening of
gravity in the limit $r \to 0$.
The above findings allow first conclusions. The RG running
of $G(r)$ in the regime where $r\gg
r_{\rm cl}$ has little quantitative influence on the gravitational
radius $r_s$. Interestingly, the precise RG running in the deep
short distance regime $r\ll r_{\rm cl}$ is also largely irrelevant for
the RG improved gravitational radius. Instead, the behaviour of
$G(r)$ and its gradient $r\,\partial_r\,G(r)$ in the regime between
$r\approx r_{\rm cl}$ and $r\approx r_s$ is mostly responsible for
the quantitative shift from $r_{\rm cl}$ to $r_s$. In consequence,
the slight differences in the distance functions \eq{distance}, \eq{linear},
\eq{D-UV} and \eq{D-interpol} are attributed to a slight variation
in the underlying RG running of $G(r)$. The RG results
from \cite{Fischer:2006fz} favour moderate values for $\gamma$, as do
regularity and minimum sensitivity considerations (see
Sec.~\ref{curvature}). In all cases studied here, the qualitative
behaviour of the horizon structure remains unchanged.
\subsection{Critical mass}\label{CM}
A direct consequence of our results from Sec.~\ref{horizons} is the
appearance of a lower bound on the black hole mass below which the RG
improved spacetime ceases to have a horizon, see \eq{Oc}. The critical
mass $M_c$ is defined implicitly via the simultaneous vanishing of
$f(r_s(M_c),M_c)=0$ and $f'(r_s(M_c),M_c)=0$ (here a prime denotes a
derivative with respect to the first argument). Using the solution $r_s(M)$ of
\eq{rs_Implicit}, we conclude that
\begin{eqnarray}\label{Mc_Implicit}
(d-3)G(r_c,M_c)&=&r_c\,G'(r_c,M_c)\,,\\
r_c&=&r_s(M_c)\,,
\end{eqnarray}
which serves as a definition for $M_c$. The classical limit is
achieved for $M_c\to 0$. If $G(r)$ is a
monotonically increasing function of $r$, we have $r\partial_r G(r)\ge 0$.
Then, away from the classical limit, there exists a unique solution
$M_c>0$ to \eq{Mc_Implicit}. Consequently, the critical mass
$M_c$ is related to the fundamental
Planck scale $M_D$ as
\begin{equation}\label{lower-1}
M_c=\zeta_c\,M_D\,.
\end{equation}
The coefficient $\zeta_c$ accounts for the renormalisation group
improvement of the black hole metric, and hence encodes the RG
effects. In the approximation \eq{G(k)}, \eq{nonlinear}, it reads
\begin{equation}
\zeta_c= \left(\frac{\tilde{\omega}}{\Omega_c}\right)^{\frac{d-3}{d-2}}
\end{equation}
where $\tilde{\omega}=\omega (\xi/c_\gamma)^{d-2}$.
The link between the RG parameters and the critical mass in units of
the fundamental Planck mass $M_c/M_D$ is displayed
in Fig.~\ref{logomega}.
The location and the number of the horizons depends explicitly on the
value of $\Omega$, which becomes
\begin{equation}\label{OmegaMc}
\Omega = \Omega_c \left(\frac{M_c}{M}\right)^{\frac{d-2}{d-3}}
\end{equation}
in terms of $M_c$, see \eq{sol-RG-3}. Therefore, below, we display our
results in terms of $M_c$. We return to the discussion of $M_c$ in
Sec.~\ref{DPS}.
\subsection{Horizons revisited}
Next, we return to the quantitative analysis of improved
metrics and present our numerical results for the improved
Schwarzschild radius.
Figs.~\ref{radius}~\ref{radius-compare} show how the Schwarzschild radius
$r_s$ depends on the mass of the black hole $M$ in various dimensions
using \eq{linear}. In these plots we considered the scenario where the
critical mass $M_c$ is equal to the Planck mass $M_D$. The suppression
is less pronounced with increasing dimension (Fig.~\ref{radius}). Also, the
deviation from classical behaviour sets in at lower masses in lower dimensions, see
Fig.~\ref{radius2}.
Next we consider varying the value of $M_c$ in units of $M_D$ while keeping
the dimensionality fixed, see Fig.~\ref{logm2}. The dashed curve
corresponds to the classical result. Depending on $M_c$, quantum
corrected curves start deviating visibly as soon as the black hole
mass is only a few $M_c$ or lower. At fixed $M/M_D$, the deviation
from classical behaviour sets in earlier for larger $M_c$.
\begin{figure}[t]
\includegraphics[width=.95\hsize]{LogomegaOct.eps}
\caption{\label{logomega} Map between the renormalisation group parameter $\tilde{\omega}$,
the critical mass $M_c$, and the Planck mass $M_D$, based on \eq{G(k)} and
\eq{linear} for various dimensions. From top to bottom:
$d=4,5,\cdots,10$. }
\end{figure}
\begin{figure}[t]
\includegraphics[width=.95\hsize]{SchwRadiusOct.eps}
\caption{\label{radius}Mass dependence of the improved
Schwarzschild radius $r_{s}(M)$ compared to its classical value
$r_{\rm cl}(M)$; $M_c=M_D$. From bottom to top: $d=4,5,\cdots,10$. The
end points denote the critical radii $r_c$.}
\end{figure}
\begin{figure*}
\subfigure[\label{radius2}\ Dimension and mass dependence of the horizon.]{\includegraphics[width=.45\hsize]{SchwRadiusDec2.eps}}
\hskip0.02\hsize
\subfigure[\label{logm2} \ $M_c$ dependence of Schwarzschild radius.]
{\includegraphics[width=.45\hsize]{Logm2Oct.eps}}
\caption{\label{radius-compare}Dependence of the renormalisation
group improved Schwarzschild radius $r_{s}(M)$
on space-time dimension and critical mass $M_c$.
End points of curves
denote the critical radius $r_c$ and dashed curves the respective classical result.
(a) $M_c=M_D$ and $d=5, 6,\cdots10$, from top right to left.
(b) $M_c/M_D=0.1,0.5,1,5,10,50$
from left to right.}
\end{figure*}
The horizon is slightly sensitive to the distance function \eq{match}, or
equivalently, to the parameter $\gamma$. Here, $\gamma$ parametrises
how rapidly $G(r)$ weakens in the cross-over regime at scales
$r\approx r_{\rm cl}$. This can be seen from Fig ~\ref{gammaOmega}. In
the decoupling limit $\gamma \rightarrow \infty$, the critical radius
$x_c \rightarrow1$ reaches the classical value. In this
limit, gravity is switched off below $r_{\rm cl}$, implying that the
Schwarzschild radius remains unchanged. For $\gamma=1$,
instead, the outer horizon $x_{+}$ decreases
rapidly as $\Omega$ increases towards $\Omega_c$.
In conclusion, the quantitative reduction of $r_s/r_{\rm cl}$ by
quantum effects can be associated to the behaviour of the running
coupling $G(r)$ and its decrease $r G'(r)$ at length scales set by the
horizon $r\approx r_s$. This decrease, in turn, can be understood via
the parameter $\gamma$ which controls how quickly quantum effects are
turned on as $r /r_{\rm cl}$ becomes small. For all matchings $k(r)\propto1/D(r)$
discussed in this paper, and for dimensionality $d\geq4$, the same
qualitative behaviour is observed. In particular the RG improvement
indicates that quantum black holes display a lower
bound \eq{lower-1} of the order of the Planck scale.
\subsection{Perturbation theory}\label{pt}
In the limit $M_D/M \ll 1$, quantum corrections become perturbative
and we can perform a large mass expansion. In particular, for
asymptotically heavy black holes we find $x_+ \rightarrow 1$, as can
be seen in Fig.~\ref{logm2} where $r_s$ approaches its classical value
for large $M$ and the dimensionless inner horizon $x_{-} \rightarrow
0$ in the large-mass limit. We note that the parameter $\Omega$ scales
as $\sim M^{-\frac{d-2}{d-3}}$. Hence a large mass expansion
corresponds to an expansion in $\Omega \ll 1$. In general, and
independently of the RG running and the matching, we find
\begin{equation}
\label{a}
x_\pm=\sum_{n=0}^\infty a^\pm_n\,\Omega^n
\end{equation}
with dimensionless coefficients $a_n$, where $a^+_0=1$ and $a^-_0=0$. The
expansion converges rapidly, see Fig.~\ref{ab}. Explicitly, the first
few coefficients read
\begin{eqnarray}
\label{linearx+}
x_{+}&=&1-\frac{1}{d-3} \Omega - \frac{d-2}{2(d-3)^2} \Omega^2
\nonumber\\ &&
- \frac{(d-1)(d-2)}{3(d-3)^3} \Omega^3+O(\Omega^4)\quad\quad
\\
\label{linearx-}
x_{-}&=& \Omega +\Omega^{d-2}+(d-2) \Omega^{(2d-5)} + \cdots\quad\quad
\end{eqnarray}
using the matching \eq{linear} and \eq{sol-RG-1}. The leading order
quantum effects modify the Schwarzschild radius $r_s=x_+\, r_{\rm cl}$
and the Cauchy horizon $r_{w}=x_{-}\, r_{\rm cl}$ as
\begin{eqnarray} \label{largemhorizon}
r_s&=& r_{\rm cl}- \frac{\Omega_c}{d-3}
\left(\frac{M_c}{M_D}\right)^{\frac{d-2}{d-3}} \frac{1}{M} + {\rm
subleading} \,,\quad\quad\quad \\ r_{w}&=&\Omega_c
\left(\frac{M_c}{M_D}\right)^{\frac{d-2}{d-3}} \frac{1}{M} + {\rm
subleading.}
\end{eqnarray}
Thus, in the limit $M_D/M \rightarrow 0$ we confirm $r_s \rightarrow
r_{\rm cl}$ and $r_w \rightarrow 0$, as expected.
Interestingly the inner horizon behaves differently if we employ the
non-linear matching \eq{D-interpol}. To that end, we again solve the
horizon condition, now given by \eq{horizoninterpol}, and expand in
$\Omega \ll 1$ to find $x_+$ and $x_-$ to leading order in $\Omega$,
\begin{eqnarray}
x_{+}&=&1- \frac{(1+\epsilon)^{\alpha-d+2}}{d-3} \Omega + {\rm
subleading}\quad\quad \\ x_{-}&=& \Omega^{\frac{1}{3-d+\alpha}}+
{\rm subleading}
\end{eqnarray}
where $\alpha$ and $\epsilon$ are given by \eq{nonalpha2} and
\eq{nonepsilon}. We note that if we take $\alpha= d-2$ we recover
\eq{linearx+} and \eq{linearx-}. In the non-linear case the outer
horizon $r_s$ has a large mass expansion similar to
\eq{largemhorizon}, whereas the inner horizon has a large mass
expansion whose leading term is proportional to a positive
power of the mass,
\begin{eqnarray}
r_s&=& r_{\rm cl}- \frac{\Omega_c (1+\epsilon)^{\alpha-d+2}}{d-3}
\left(\frac{M_c}{M_D}\right)^{\frac{d-2}{d-3}}
\frac{1}{M}\quad\quad\quad \\ r_{w}&=&\0{1}{M_D} \left(\frac{M_c
\Omega_c}{M_D}\right)^{\rho_0} \left(\frac{M}{M_D}\right)^{\rho_1}
\end{eqnarray}
plus terms subleading in $M$. We have introduced $\rho_0=
\frac{d-2}{(d-3)(\alpha +3 -d)}$,
$\rho_1=\frac{5-2d+\alpha}{(d-3)(\alpha+3-d)}$ and
$\rho_1+\rho_2=\01{d-3}$. For $d=4$, $\rho_{1}=0$ for $d>4$ we find
$1>\rho_1>0$. This implies that in the limit $M \rightarrow \infty$
the Cauchy horizon $r_w$ will approach a constant for $d=4$. In higher
dimensions $d>4$, $r_w$ will increase with mass as $r_{w} \sim
M^{\rho_1}$, whereas the ratio $r_{w}/r_{s} \sim M^{2-d} \,
\rightarrow 0$ in the large mass limit.
\subsection{Threshold effects}\label{te}
The RG improved black hole displays an interesting threshold behaviour
in the vicinity of $M\to M_c$. This can be understood as
follows. Suppose we read \eq{f-RG} as a function of $r$ and $M$,
$f(r,M)$, and perform a Taylor expansion in both variables. The
outcome is then evaluated at the horizon $r=r_s(M)$ where
$f(r_s(M),M)=0$. Independently of the chosen expansion point
$(r_0,M_0)$ with $r_0=r_s(M_0)$, we find
\begin{eqnarray}\nonumber
0&=&
\sum_{n=1}\left[ \frac{1}{n!}(M-M_0) ^n\frac{\partial^n f}{\partial
M^n}
\label{threshold}
+\frac{1}{n!}(r-r_0)^n\frac{\partial^n f}{\partial r^n}\right]
\end{eqnarray}
at the horizon. Note that the derivatives are evaluated at
$(r,M)=(r_0,M_0)$. If the RG running of $G$ is $M$-independent, the
expansion has only a linear term in $(M-M_0)$. At threshold where
$r_0=r_c$, we furthermore have $\partial f/\partial r|_{r_c} = 0$. In
addition, $\partial f/\partial M|_0$ is always non-zero. Therefore,
close to $M\approx M_c$, $f(r(M),M)=0$ can only be satisfied if
\begin{equation}\label{sqrt}
r_s(M)-r_c\sim \sqrt{M-M_c}\,,
\end{equation}
provided that $\partial^2 f/\partial r^2|_{r_c} \neq 0$. More generally, if the
first non-vanishing derivative $\partial^n f/\partial r^n|_{r_c}$
occurs at order $n$, the threshold behaviour \eq{sqrt} becomes
\begin{equation}
\label{higher} r_s(M)-r_c\sim \sqrt[n]{M-M_c}\,.
\end{equation}
The generic case encountered in this paper, for all RG runnings employed, is
$n=2$. Consequently, at threshold, we have the expansions
\begin{equation}
\label{b}
x_\pm=\sum_{n=1}^\infty b^\pm_n\,
\left(\frac{M}{M_c}-1\right)^{n/2}
\end{equation}
with dimensionless coefficients $b^\pm_n$. {This is equivalent to an
expansion in powers of $\sqrt{\Omega_c-\Omega}$.} This expansion
converges rapidly as can be seen from Fig.~\ref{ab}, where the first
few terms (up to $n=6$) are enough to match the full solution even for
small $\Omega$.
Explicitly, using the matching \eq{linear}, the behaviour \eq{sqrt}
reads to leading order
\begin{equation}
r_s(M)-r_c =
\frac{(G_0 M_c)^{1/(d-3)}}{ \sqrt{\012(d-3)M_c}} \sqrt{M-M_c}\,.
\end{equation}
In the light of the above, the dependence of the horizon radius on the
black hole mass can easily be understood, see Fig.~\ref{radius}. For
large black hole mass $M\gg M_c$, $\partial_r f$ is non-vanishing at
$f(r_s)=0$, implying that the linear terms in \eq{threshold} have to
cancel. This leads to the very soft dependence of $r_s/r_{\rm cl}$ on
$M_c/M$ for large $M$. With decreasing $M$, $\partial_r f$ is
decreasing as well, thereby increasing the admixture from $(r-r_0)^2$
corrections. The latter fully take over in the limit $M\to M_c$,
leading to non-analytical behaviour \eq{sqrt} which is nicely seen in
Fig.~\ref{radius}. We stress that this structure is independent of the
dimension as long as $d>3$.
\begin{figure*}
\subfigure[\ Perturbative approximation of the horizon.]
{\includegraphics[width=.45\hsize]{pD8Expansion.eps}}
\hskip0.09\hsize
\subfigure[\ Threshold approximation of the horizon.]
{\includegraphics[width=.45\hsize]{pD8ExpansionW.eps}}
\caption{\label{ab} Location of the event horizon $x_+(\Omega)\equiv r_s/r_{\rm cl}$
as a function of the parameter $\Omega$ in $d=8$ (thick line)
within various approximations (thin lines). (a)
Perturbative expansion about $\Omega=0$ using \eq{a} at order $n=1,2,\cdots,20$,
approaching the exact solution adiabatically (from top to bottom).
(b) Threshold expansion about the critical point $\Omega=\Omega_c$ using \eq{b} at order $n=1,2,\cdots,9$,
alternating towards the full solution.}
\end{figure*}
\subsection{Temperature and specific heat}\label{t}
The threshold behaviour of Sec.~\ref{te} has important implications
for the thermodynamics of black holes. Based on the RG improvement
used here, the Bekenstein-Hawking temperature is given by \cite{Hawking:1979,Gibbons:1976ue}
\begin{equation}
T=\frac{f'(r_s(M),M)}{4\pi}\,,
\end{equation}
where the prime denotes a derivative with respect to the first
argument. We conclude that the temperature is bounded, $T(M)\le T_{\rm cl}(M)$. Equality is
achieved for asymptotically large black hole mass, where the
temperature scales inversely with mass, $T\sim 1/M$, and the specific
heat
\begin{equation}
C=\frac{\partial M}{\partial T}
\end{equation}
is negative. In this regime, it is expected that the black hole evaporates
through the emission of Hawking radiation, thereby lowering its mass
through the influx of negative energy.
At threshold with $M\to M_c$, however, the black hole
temperature vanishes because $f'(r_c)\to 0$. In the vicinity of
$M\approx M_c$, the temperature of the black hole scales
non-analytically,
\begin{equation}
\frac{\partial T}{\partial M}\sim
\frac{\partial r_s(M)}{\partial M}\sim \frac{1}{\sqrt{M-M_c}}>0
\end{equation}
as a consequence of \eq{sqrt}. Therefore, the specific heat has become
positive, and the temperature displays a maximum $T(M)\le T_{\rm max}$
for all $M$. Furthermore, since $\partial T/\partial M$ vanishes at
$T_{\rm max}$, the specific heat changes its sign through a pole. We
conjecture that black holes with mass $M_c$ constitute cold remnants,
provided they can be reached through an evaporation process. This would
require, however, that $T_{\rm max}$ remains below the fundamental Planck scale.
\subsection{Renormalisation and the Planck scale}\label{DPS}
We summarise our results.
The main physics of this paper originates from a
new mass scale $M_c$, which is absent in the classical theory.
Its existence is due to
quantum gravity corrections, implemented on the level of the
metric.
For black hole mass $M$ large compared to $M_c$, renormalisation group
corrections to the metric are small. The
gravitational force remains strong enough to allow for black
holes. Improved black hole metrics display horizons of the
order of the classical horizon, the specific heat stays negative and the temperature
scale inversely proportional to mass, modulo small quantum corrections. This is the
semi-classical regime of the theory.
Once the mass $M$ approaches $M_c$, we observe the transition from
strong to weak gravity. In its vicinity, renormalisation effects become of
order one, the reduction of the event horizon becomes more pronounced,
the specific heat has become positive and the
thermodynamical properties are no longer semi-classical. This is the
Planckian (or quantum) regime of the theory.
When $M$ drops below $M_c$, renormalisation
group corrections to the metric have become strong. The gravitational force
has weakened significantly, to the point that improved black hole space-times
no longer display a horizon. This is the deep UV scaling regime of the
theory. The improved metric differs qualitatively from the classical one. Hence, the
applicability of our renormalisation group improvement
becomes doubtful, and conclusions from this regime
have to be taken with care.
If the renormalisation
effects of the black hole space-time are parametrically
strong, $\tilde\omega\gg 1$,
the scale $M_c$ grows large, and parametrically larger than the
Planck scale $M_D$. In turn, for weak renormalisation
$\tilde\omega\ll 1$, the scale $M_c$ remains small as well.
We note that $M_c$ vanishes in the classical limit where
quantum corrections are switched off. The reason for this is
that the Schwarzschild
solution of classical general relativity does not predict its own
limit of validity under quantum corrections. Interestingly, the underlying fixed point is
not primarily responsible for the existence of the lower bound $M_c$. Other
ultraviolet completions of gravity such as string theory, loop quantum gravity or
non-commutative geometry can lead a similar weakening of the gravitational force at
length scales of the order of the Planck length.
To conclude, the improved metric changes qualitatively at $M\approx
M_c$. Therefore it is tempting to interpret $M_c$ as a `renormalised' Planck
scale. Its numerical value depends on the precise renormalisation group running.
As long as the latter is driven by the
gravitational self-coupling only, it is natural to have $M_c$ of
the order of $M_D$. This may be different once strong renormalisation effects are
induced by external mechanisms, eg.~through the coupling to a large number
of matter fields.
\section{Space-time structure and Penrose diagram}\label{STS}
In this section, we study the implications of quantum gravitational
effects on the space-time structure of black holes, including a
discussion of critical black holes, an analogy with
Reissner-Nordstr\"om black holes, an interpretation in terms of an
effective energy momentum tensor, the (absence of) curvature
singularities at the origin, and the causality structure and Penrose
diagram of quantum black holes.
\subsection{Critical black holes}
The space-time structure of a critical black hole with mass $M=M_c$
has a single horizon at $r_c=r_{\rm cl}\,x_c$ and $x_{-}=x_{+}=x_c$
where the function $f(x)$ has a double zero $f(x_c)=0$. For the
matching \eq{linear} $x_c$ is given explicitly by \eq{xcr}. The
near-horizon geometry of a critical black hole is obtained by
expanding $f(x)$ around $x=x_c$. We find
\begin{equation}
f(x)= \0{1}{2} \, \bar{x}^2\, f''(x_c,\Omega_c)
\end{equation}
where $\bar{x}\equiv x-x_c$ and the double prime represents a second
derivative with respect to $x$. Therefore, we can write the line
element in terms of the coordinate $\bar{r}=r-r_c$ as
\begin{eqnarray}
\label{AdS}
ds^2&=&
-\frac{\bar{r}^2}{r_{\rm AdS}^2}dt^2
+ \frac{r_{\rm AdS}^2}{\bar{r}^2} dr^2
+ r_c^2\, d\Omega^2_{d-2}\quad
\end{eqnarray}
The metric \eq{AdS} is the product of a two-dimensional anti-de Sitter
space with a $(d-2)$-sphere, ${\rm AdS}_2 \times S^{d-2}$, and depends
on their respective radii
\begin{eqnarray}
\label{rAdS}
r_{\rm AdS}&=&(G_0 M_{\rm AdS})^{{1}/{(d-3)}}\\
r_c&=& x_c\,(G_0\,M_c)^{1/(d-3)}
\end{eqnarray}
The curvature of the anti-de Sitter part is determined by the mass
parameter $M_{\rm AdS}$,
\begin{equation}\label{AdSmass}
M_{\rm AdS}= M_c \left(\frac{1}{2}f''(x_c, \Omega_c)\right)^{-\012(d-3)}\,.
\end{equation}
Using \eq{linear} for $d=4$ we have $M_{\rm AdS}=
\frac{1}{\sqrt{2}}M_c$. For higher dimensions $d=6,8, 10$ we find
$M_{\rm AdS}/M_c\approx 0.14, 0.017, 0.0016$, respectively. For all
dimensions, we have $M_{\rm AdS}<M_c$. We note that the metric
\eq{AdS} is of the form of a Robinson-Bertotti metric for a constant
electric field.
\subsection{Reissner-Nordstr\"om-type metrics}
It is interesting to compare the RG improved black hole with the
well-known Reissner-Nordstr\"om solution of a charged black
hole in higher dimensions \cite{Myers:1986un}. The latter is defined
via the lapse function
\begin{equation}\label{fRN}
f_{\rm RN}(r)=1-\frac{G_0 M}{r^{d-3}} + \frac{G_0 e^2}{r^{2(d-3)}}
\end{equation}
where $e^2$ denotes the charge of the black hole (squared). The charge
has the mass dimension $[e^2]=4-d$. The physics of the
Reissner-Nordstr\"om black hole is best understood in terms of the
dimensionless parameter
\begin{equation}\label{oRN}
\Omega_{\rm RN}=\frac{e^2}{G_0 M^2}
\end{equation}
which measures the relative strength of the competing terms on the rhs
of \eq{fRN}. In terms of \eq{oRN} and using $x=r/r_{\rm cl}$ , the
lapse function becomes
\begin{equation}\label{RN}
f_{\rm RN}(x)=1-\frac{1}{x^{d-3}}+\frac{\Omega_{\rm RN}}{x^{2d-6}}\,.
\end{equation}
For $\Omega_{\rm RN}>\frac{1}{4}$ the spacetime has no horizons and
exhibits a naked singularity. For $\Omega_{\rm RN}<\frac{1}{4}$ the
spacetime displays two horizons, whereas for $\Omega_{\rm RN} =
\frac{1}{4}$ the black hole displays a single horizon. Therefore
$\Omega_{\rm RN}=\frac{1}{4}$ is referred to as a extremal black hole
with critical mass $M_{\rm RN, c}=2\sqrt{e^2/G_0}$. The radius of the
extremal black hole is given by $r_{\rm RN,c}= 2^{-1/(d-3)}\, r_{\rm
cl}$.
Reissner-Nordstr\"om spacetimes share some of the qualitative features
of RG improved higher dimensional black holes discussed in this paper.
If we consider a quantum black hole using the matching \eq{linear} and
expand the lapse function to leading order in $\Omega$, we find
\begin{equation}\label{LO}
f_{\rm LO}(x)=1-\frac{1}{x^{d-3}}+\frac{\Omega}{x^{2d-5}}+{\rm subleading}\,.
\end{equation}
In either case \eq{RN} and \eq{LO}, the relevant physics originates
from competing effects: a leading order Schwarzschild term
$~\,-1/r^{d-3}$, which is counterbalanced by either the charge,
parametrised by $\Omega_{RN}\sim e^2$, or by quantum corrections due
to a running gravitational coupling, parametrised by
$\Omega\sim\tilde{\omega}$. The correction terms become quantitatively
dominant with decreasing $r \rightarrow 0$. We note that \eq{RN} and
\eq{LO} are formally equal for $\Omega_{\rm RN}=\Omega \, x$. In
either case, in the large mass limit $M \rightarrow \infty$ we find an
outer horizon $f(x_{+})=0$ for $x \approx 1$. It follows that the near
horizon geometry of a quantum black hole is approximately that of a
Reissner-Nordstr\"om black hole of charge $e^2=\tilde{\omega}\,
r_{\rm cl}^{d-4}$, and in the large mass limit.
Next we consider the near horizon geometry of an extremal
Reissner-Nordstr\"om black hole, which is of the ${\rm AdS}_2 \times
S^{d-2}$ type. The line element is given by \eq{AdS} where
\begin{eqnarray}
\label{AdSradiusRN}
r_c&=&\left(\s012 G_0 M_{\rm RN, c}\right)^{1/(d-3)}\\
\label{AdSmassRN}
M_{\rm AdS}&=& M_{RN, c}\, \left(\s0{1}{2}f_{RN}''(x_c, \Omega_c)\right)^{{-(d-3)}/{2}}\,.\quad
\end{eqnarray}
For $d=4$ we find $M_{\rm AdS}= \frac{1}{2}M_{\rm RN, c}$. For higher
dimensions $d=6, 8$ and $10$, we obtain $M_{\rm AdS}/M_{\rm RN, c}
\approx 0.02, 0.0002$ and $ 6 \times 10^{-7}$, respectively. The
decreasing of $M_{\rm AdS}/M_{\rm RN, c} $ with dimension is similar
to the decreasing of $M_{\rm AdS}/M_{c}$ for the critical black hole
\eq{AdSmass}.
\subsection{Effective energy-momentum tensor}
In this paper we have obtained our results by replacing the classical
value of Newton's constant $G_0$ with a running constant $G(r)$. It
is interesting to ask whether this modification could have arisen from
an explicit source term, the energy-momentum tensor, for Einstein's
equations. The answer is affirmative, and obtained by inserting the RG
improved metric into the left hand side of the Einstein equations
$G^{\mu \nu}= 8 \pi\, G_0\, T^{\mu \nu}$. The non-vanishing components
are the diagonal ones $T^\mu_{\, \nu}= {\rm diag}(-\rho, p_r,
p_\perp,..,p_\perp)$, given by
\begin{equation} \rho=-p_r=\frac{G'(r)\,M_{\rm
phys}}{S_{d-2}\, G_0\, r^{d-2}}
\end{equation}
\begin{equation}
p_\perp=-\frac{G''(r)\,M_{\rm phys}}{(d-2) S_{d-2} G_0 r^{d-3}}
\end{equation}
where $S_{d-2}={2 \pi^{(d-1)/2}}/{\Gamma((d-1)/2)}$. Integrating the
energy density $\rho$ over a volume of radius $r$ one finds the
effective energy within that radius,
\begin{equation} E(r)=S_{d-2} \int^r_0 dr'
\rho(r')r'^{d-2}= \frac{G(r) \,M_{\rm phys}}{G_0} \, ,
\end{equation}
where we assume $G(r)$ obeys the limits $G(r)\rightarrow 0$ for
$r\rightarrow0$ and $G\rightarrow G_0$ for $r\rightarrow \infty$. As
such we note that $E(\infty)=M_{\rm phys}$ the physical mass. We also
make the observation that replacing $G(r)\,M_{\rm phys} \rightarrow
G_{0}E(r)$ leaves the metric invariant.
\subsection{Absence of curvature singularities} \label{curvature}
In this section, we discuss the $r\to 0$ limit of asymptotically safe
black holes and the absence of curvature singularities. Classical
Schwarzschild solutions display a coordinate singularity at $r=r_{\rm
cl}$ where $f(r)^{-1} \rightarrow \infty$. Curvature invariants
remain well-defined and finite at the horizon, which shows that the
singularity is only an apparent one.
A curvature singularity in the classical metric is found at $r\to 0$,
where $f(r)\rightarrow -\infty$ and the Ricci scalar diverges as $R
\sim r^{1-d}$. This curvature singularity implies the break-down of
classical physics at the centre of a black hole. It is expected that
quantum fluctuations should lead to a less singular or finite
behaviour as $r \rightarrow 0$.
Within the renormalisation group
set-up studied here, the main new input is the anti-screening of the
gravitational coupling.
Consequently, $G(r)/G_N$ becomes very small, thereby modifying the
$r\to 0$ limit. For small $r/r_{\rm cl}\ll 1$, we have
\begin{equation}\label{smallr}
f(r) = 1-(\mu\,r)^\sigma\,+\,{\rm subleading}
\end{equation}
where the mass scale $\mu$ and the parameter $\sigma$ are fixed by the
renormalisation group and the matching condition discussed in
Sec.~\ref{QG}. We note that a value of $\sigma=2$ would correspond to
a de Sitter core with cosmological constant
\begin{equation}\label{LdS}
\Lambda_{\rm dS}=\frac{1}{2}(d-1)(d-2)\mu^2\,.
\end{equation}
More generally, the value of $\sigma$ depends on the detailed short
distance behaviour. We find $\sigma_{\rm ir}=1$ using \eq{linear} for
all values of $d\geq4$, and $\sigma_{\rm uv}=\s012(d^2-5d+8)$ using
\eq{D-UV}. For $d\geq4$ the latter takes values $\sigma_{\rm
uv}\geq2$. In contrast to this, the classical solution displays
$\sigma_{\rm cl}= 3-d$. Using the matching \eq{nonlinear} with
parameter $\gamma$ we have $\sigma=\gamma(d-2)-d +3$. Consequently, a
de Sitter core is achieved for
\begin{equation}
\gamma_{\rm dS}=(d-1)/(d-2)\end{equation}
in the limit $r \rightarrow 0$.
Next, we calculate the Ricci scalar, the Riemann tensor squared and
the Weyl tensor squared in the limit $r\rightarrow0$, using
\eq{smallr}. The results are
\begin{eqnarray}\nonumber
R&=& F_R\cdot(\mu\, r)^{\sigma-2}\,\mu^2\\
R^{\mu\nu\kappa\lambda}R_{\mu\nu\kappa\lambda}&=&F_{\rm Riem}\cdot(\mu\,
r)^{2\sigma-4}\,\mu^{4}\nonumber \\
C^{\mu\nu\kappa\lambda}C_{\mu\nu\kappa\lambda}&=&F_C\cdot(\mu\,
r)^{2\sigma-4}\,\mu^{4}
\end{eqnarray}
modulo subleading corrections. The coefficients are
\begin{eqnarray}
F_R&=& (\sigma+d-2)(\sigma+d-3)\nonumber \\
F_{\rm Riem}&=&\sigma^4-2\sigma^3+(2d-3)\sigma^2+2(d-2)(d-3)\nonumber\\
F_C&=&\frac{d-3}{d-1}(\sigma-1)^2(\sigma-2)^2
\end{eqnarray}
Clearly, the curvature singularity is absent as soon as $\sigma\ge 2$,
which in general is achieved for the matchings employed here
including \eq{D-UV}. For the matching \eq{linear}, however, we have
$\sigma=1$ and conclude that in this case the remaining curvature
singularity reads $R \sim \frac{1}{r}$. This is still a significant
reduction in comparison with the behaviour $\sim r^{1-d}$ within the
classical Schwarzschild solution, and indicates that the weakening of
gravitational interactions leads to a better short distance behaviour.
The Riemann-squared coefficient is non-zero for all values of $\sigma$
when $d\geq4$. The Weyl-squared term has the same $r$-dependence as
the Riemann -squared term, but its coefficient vanishes for both
$\sigma=1$ and $\sigma=2$. Hence, there is no choice for $\sigma$
which makes all three coefficients vanishing.
We are lead to the following conclusions. Regularity of an asymptotically safe
black hole requires $\sigma\geq 2$. The RG study indicates that the behaviour
for the physical theory lies in between the limits set by $\sigma_{\rm ir}
\le \sigma_{\rm phys}\le\sigma_{\rm uv}$. It is tempting to speculate that the
physical value would read $\sigma=2$ corresponding to a de Sitter core
with positive cosmological constant set by \eq{LdS}. A distance function
with effective index $\gamma\ge\gamma_{\rm dS}$ together with
a momentum-scale RG for Newton's coupling provides for a
singularity-free metric for all $r$. This is a very mild constraint on the RG
running, as $\gamma_{\rm dS}\in[1,\frac{3}{2}]$ is very close
to $\gamma_{\rm ir}=1$ for all $d\ge4$.
As a last observation, we treat $\gamma$ as a free parameter and
employ a principle of minimum sensitivity (PMS) condition to identify
its best match value $\gamma_{\rm PMS}$ \cite{Stevenson:1981vj}. Since
the horizon radius $r_s=r_s(\gamma)$ at fixed black hole mass depends
monotonically on $\gamma$, a PMS condition singles out the boundaries
of the parameter space given by the decoupling limit $1/\gamma\to 0$,
and the de Sitter limit $\gamma=\gamma_{\rm dS}$. The decoupling limit
corresponds to the switching-off of gravity. Hence, we conclude that a
minimum sensitivity condition singles out $\gamma_{\rm
PMS}=\gamma_{\rm dS}$.
\subsection{Kruskal-Szekeres coordinates}
In this section we introduce Kruskal-Szekeres coordinates which remove
the coordinate singularities at the horizons. This is the first step
towards a discussion of the causal structure of asymptotically safe black holes
and their Penrose diagrams.
Here we consider the case where $M>M_c$
such that the spacetime has two horizons; the outer horizon $r_s
\equiv r_{\rm cl}\,x_+$ and the Cauchy horizon $r_w \equiv r_{\rm
cl}\,x_{-}$. For simplicity we will consider the linear matching
\eq{linear} where the lapse function is given by \eq{f-L} such that
$\alpha=d-2$. The horizons are found by the real positive roots of
\eq{sol-RG-1}. In general there will be exactly $\alpha=d-2$, possibly
complex, roots. In the regime of interest where $0< \Omega<\Omega_c$,
we have always two real positive roots $x_\pm$. In even or odd
dimensions, we additionally find $(d-4)/2$ pairs of complex conjugate
roots, or a real negative root and $(d-5)/2$ pairs of complex
conjugate roots, respectively. Therefore, we decompose
\begin{eqnarray}
\Delta &\equiv& x^{\alpha}+\Omega-x \nonumber\\
&=& (x-x_{+})(x-x_-)\prod_{i=1}^{\alpha-2}(x-z_i).
\end{eqnarray}
into the two real roots $x_{\pm}>0$ and the remaining $d-4$ roots
$z_i$. In terms of these, we have
\begin{equation}
\Omega= (-1)^{\alpha}\, x_+\, x_-\, \prod_{i=1}^{\alpha-2} z_i \,.
\end{equation}
We express the line element in terms of the roots and the
dimensionless radial coordinate $x$
\begin{equation}
ds^2=-\frac{\Delta}{x^\alpha+\Omega}dt^2
+ \frac{x^\alpha+\Omega}{\Delta} dr^2 + r^2\ d\Omega^2_{d-2} \, .
\end{equation}
Next we express the line element in terms of Kruskal-Szekeres type
coordinates to remove the coordinate singularities. We will follow the
method as outlined in \cite{Townsend:1997ku} for a Reissner-Nordstr\"om black
hole with two horizons. First we define the dimensionless tortoise
coordinate
\begin{equation}
dx^{*}=\frac{x^\alpha+\Omega}{\Delta}\,dx
\end{equation}
It is then clear that radial null geodesics correspond to $t/r_{\rm
cl} = \pm x^*$. Performing the integral we find
\begin{eqnarray}
x^*&=&x+\frac{1}{2\kappa_+} \ln( \left|x-x_+\right|) + \frac{1}{2\kappa_-} \ln(\left|x-x_-\right|)
\nonumber\\ &&
\label{xstar}
+ \sum^{\alpha-2}_{i=1} \frac{1}{2\kappa_i} \ln((x-z_i))+ {\rm constant}\\
\kappa_i&=&\frac{(z_i-x_+)(z_i-x_-)\prod^{\alpha-2}_{j \neq i} (z_i-z_j)}{2z_i}\quad \ \ \ \\
\kappa_+&=&\frac{(x_+-x_-)\prod^{\alpha-2}_{j=1} (x_+-z_j)}{2x_+}\\
\kappa_-&=&\frac{(x_{-}-x_+)\prod^{\alpha-2}_{j=1} (x_--z_j)}{2x_-}
\end{eqnarray}
We now introduce advanced and retarded time coordinates given by
\begin{eqnarray}
v=x^*+w\\
u=x^*-w
\end{eqnarray}
where $w$ is the dimensionless time $w \equiv t/r_{\rm cl}$. We then
define the coordinates
\begin{eqnarray}\label{KS}
V^{\pm}=e^{\kappa_\pm\,v}\\
U^{\pm}=-e^{\kappa_\pm\,u}
\end{eqnarray}
These are the KS-type coordinates for quantum black holes. The product
\begin{equation} U^\pm V^\pm= -e^{2\kappa_\pm x^*} \end{equation} is a constant for any
given radius $x$. In terms of the coordinates $U^+$ and $V^+$ the
line element becomes \begin{eqnarray} ds^2=&-&\left(\frac{r_{\rm
cl}}{\kappa_{+}}\right)^2 \, e^{-2 \kappa_{+}\,x^*}
\frac{\Delta}{x^\alpha+\Omega} dU^+dV^+ \quad \ \ \ \nonumber\\ &+&
r^2\ d\Omega^2_{d-2} \end{eqnarray} Inserting $x^*$ given by \eq{xstar} we find
\begin{eqnarray}
ds^2&=&-r_{\rm cl}^2 \, F_+(x)\, dU^+\, dV^+ + r^2\ d\Omega^2_{d-2}
\\
F_{+}&=& \frac{e^{-\kappa_+x}}{x^\alpha+\Omega}\frac{\kappa_+^{-2}(x-x_-)}{(x-x_-)^{\frac{\kappa_+}{\kappa_-}}} \prod^{\alpha-2}_{i=1} \frac{x-z_i}{(x-z_i)^{\frac{\kappa_+}{\kappa_i}}}\,.\quad\quad
\end{eqnarray}
The singularity in the $x_+$ coordinate has been removed and the
metric covers regions of space time for $x>x_-$. There remains a
singularity at $x=x_-$, and, therefore, the metric does not cover the
region $x\leq x_{-}$. Instead we use the coordinates $U^-$ and $V^-$
in terms of which the line element is given by
\begin{eqnarray}
ds^2&=&-r_{\rm cl}^2 \, F_-(x)\, dU^-\,dV^- + r^2\ d\Omega^2_{d-2}\\
F_{-}&=& \frac{e^{-\kappa_+x}}{x^\alpha+\Omega}
\frac{\kappa_-^{-2}(x-x_+)}{(x_+-x)^{\frac{\kappa_-}{\kappa_+}}}
\prod^{\alpha-2}_{i=1} \frac{x-z_i}{(x-z_i)^{\frac{\kappa_-}{\kappa_i}}}\,.\quad\quad
\end{eqnarray}
Hence the singularity at $x=x_{-}$ is removed in these coordinates and
the metric is well defined in the region $x<x_{+}$. The singularity
at $x=x_{+}$ remains in this parametrisation and does not cover the
region $x>x_+$. The coordinates \eq{KS} are defined such that for
ingoing null rays $V^\pm={\rm constant}$ and for outgoing null rays
$U^\pm={\rm constant}$.
\begin{figure}[t]
\includegraphics[width=1.25 \hsize]{Penrose2.eps}
\caption{\label{Penrose} The Penrose diagram of a
quantum black hole with $M>M_c$. The black curves in
regions I and V are curves of constant $t$. The
green (blue) [red] curves are curves of constant $r$
in region I and V (II and IV) [III],
respectively. In- and outgoing radial null geodesics
are at $45^o$. Curves 1., 3. and 4. correspond to
schematic plots of various solutions to the
equations of motion. The points $i^0$, $i^+$ and
$i^-$ denote spatial infinity, future infinity and
past infinity, respectively. $\mathcal{J^-}$ and
$\mathcal{J^+}$ denote past and future null infinity
(see text).}
\end{figure}
\subsection{Causality and Penrose diagram}
The global structure of the black hole can be represented by a Penrose
diagram. To produce the diagram we make an analytical continuation of
the KS-type coordinates and then map them to a finite interval
\begin{eqnarray}
V^\pm \rightarrow \tanh(V^\pm) \\
U^\pm \rightarrow \tanh(U^\pm).
\end{eqnarray}
The resulting Penrose diagram is shown in Fig.~\ref{Penrose}. The
causal structure can be understood by noting that null geodesics are
always at $45^o$ such that ingoing photons point ``north-west'' and
outgoing photons point ``north-east''. Regions I, II and III
correspond to $x>x_+$, $x_-<x<x_+$ and $x<x_-$, respectively, where
$x_+$ denotes the outer horizon and $x_-$ the inner Cauchy horizon in
units of $r_{\rm cl}$. The other regions are the analytical
continuations; in particular regions IV and V correspond to
$x_-<x<x_+$ and $x>x_+$. Surfaces of constant $r$ in region II
($x_-<x<x_+$) are trapped surfaces such that all null geodesics move
towards the inner horizon. On the other hand region IV defines a
white hole where all null geodesics point towards $r_+$.
To get an idea of the causal structure experienced by an in-falling
observer we follow the standard procedure of considering a radially
moving test particle as was done in the $d=4$ case
\cite{Bonanno:2000ep}. We define the dimensionless proper time of the
radial particle $d\tau^2=ds^2/r_{\rm cl}^2$. A constant of motion
$\zeta$ is defined by the Killing vector equation corresponding to the
time independent nature of the metric,
\begin{equation}
\zeta = f(x) \frac{dw}{d\tau} \, .
\end{equation}
From the form of the metric \eq{ds2} the equations of motion for the test particle can then be given
in terms of $\zeta$:
\begin{equation}\label{eqmotion}
\dot{x}^2= \zeta^2-f(x)
\end{equation}
(dots denote derivatives with respect to proper time $\tau$.) We
define a Newtonian-like potential, $\Phi(x)=\frac{1}{2}(f(x)-1)$, to
write an equation for the proper acceleration of the test particle
\begin{equation}
\ddot{x}=-\frac{\partial \Phi(x)}{\partial x}\, .
\end{equation}
This equation can be checked by differentiating \eq{eqmotion} with
respect to $\tau$. For the linear matching \eq{linear} the proper
acceleration is given by,
\begin{equation}
\ddot{x} =-\frac{1}{2}\frac{(d-3)x^{d-2}-\Omega}{(x^{d-2}+\Omega)^2} \, .
\end{equation}
From \eq{eqmotion} write down an ``energy'' equation:
\begin{equation}
E \equiv \frac{\zeta^2-1}{2}=\frac{1}{2}\dot{x}^2+\Phi(x) \,
\end{equation}
For different values of $E$ we analyse various solutions to the
equations of motion for radially moving test particles. The potential
takes its maximum value $\Phi_{max}=0$ at $r=0, \infty$ and, for
$M>M_{c}$, its minimum value will be $\Phi_{min}<-\frac{1}{2}$. The
different solutions discussed below are shown as curves in
Fig. \ref{Penrose}.
\begin{enumerate}
\item For $E=0$ the particle has zero velocity at $r=0$ and
$r=\infty$. For the linear matching the particle will start in
region I with a non-zero velocity and cross the horizons into
regions II and III in a finite proper time. The particle will then
reach the centre of the black hole where it feels a repulsive force
with a strength of $1/(2r_s\Omega)$. This force will bounce the
particle back into regions IV and V where it will escape to
infinity.
\item For $E>0$ the motion of the particle will be unbounded since it
has a non-zero velocity at all points in spacetime. Starting from
region I the particle will again move to the centre of the black
hole crossing both horizons in a finite time $\tau$. But at $r=0$
the particles energy will be enough to overcome the repulsive force
and will pass through the centre of the black hole into regions IV
and V where it will escape to infinity.
\item For $-0.5<E<0$ the particle starts with zero velocity in region
I and continues to move into regions II and then into III where it
has an inflection at $r>0$. Here the particle is bounced into
regions IV and V. In region V it has a second inflection point at a
radius equal to it's initial position in region I. The particle's
motion is therefore bounded moving in and out of the black hole into
different regions of spacetime.
\item For $\Phi_{min}<E<-0.5$ the particle's motion is bound to region
II in which it has two inflection points which it moves between
eternally.
\end{enumerate}
A discussion of causal structures, taking into account
the time-dependent evaporation effects of asymptotically safe black holes,
will be summarised elsewhere \cite{erg2008}.
\section{Black hole production}\label{production}
In this section, we apply our results to the
production of mini-black holes in higher-dimensional particle physics
models of TeV scale quantum gravity.
\subsection{Large extra dimensions}
The scenario of large extra dimensions assumes that gravity propagates
in $d=4+n$ dimensions, whereas matter fields are confined to a
four-dimensional brane \cite{ArkaniHamed:1998rs}. The $n$ extra dimensions are
compactified with compactification radius $L$. For simplicity, we
assume that all radii are of the same order of magnitude, which can be
relaxed if required. The presence of extra dimensions allows for a
fundamental $d$-dimensional Planck scale $M_D$ of the order of the
electroweak scale $ \sim 1$TeV. The relationship between the effective
4-dimensional Planck scale $M_{\rm Pl}$ and the $d$-dimensional Planck
scale is given by
\begin{equation}
M_{\rm Pl}^2 \approx M_D^2 (M_D \,L)^n\,.
\end{equation}
Furthermore, we require the scale-separation $M_D \,L\gg 1$ to achieve
a low fundamental quantum gravity scale. This implies that the length
scale $L$ at which the extra dimensions become visible is much larger
than the fundamental length scale $1/M_D$ at which the quantum gravity
effects become important. Consequently, at energy scales $E\approx
M_D$, the full $d$-dimensional space-time is accessible to gravity,
and our previous findings are applicable.
\subsection{Production cross section}
Here, we apply our results \cite{erg2008} to the production cross section for
mini-black holes at particle colliders. In these models, the elastic
black hole production cross section for parton-parton scattering at
trans-Planckian center-of-mass energies $\sqrt{s}\gg M_D$ is
semi-classical, provided curvature effects are small
\cite{Dimopoulos:2001hw,Giddings:2001bu,Banks:1999gd,Hsu:2002bd,Eardley:2002re}. Then, on the parton level,
the geometric cross section reads
\begin{equation}\label{sc}
\hat\sigma_{\rm cl}(s)\approx \pi\,r_{\rm cl}^2(M_{\rm phys}=\sqrt{s})\,\theta(\sqrt{s}-M_{\rm min})\,,
\end{equation}
with the physical mass replaced by the center-of-mass energy
$\sqrt{s}$. There are formation factor corrections to \eq{sc} which
have been identified in the literature, taking into account
inefficiencies in the production process (see \cite{Kanti:2004nr} for
reviews). Those have not been written out explicitly as they are
irrelevant to our reasoning. For phenomenological applications, it is
often assumed that the minimal mass $M_{\rm min}$ is of the order of a
few $M_D$, limiting the regime where the semi-classical theory is
applicable.
Our study adds two elements to the picture. The first one relates to
the threshold mass, indicating that $M_{\rm min}$ may in fact be
lower, possibly as low as the renormalised Planck mass
\begin{equation} M_{\rm min}=M_c\,.
\end{equation}
This is a direct consequence of
the RG running of the gravitational coupling, with $M_c$ relating to
the critical physical mass (defined as in \eq{M-def}), thereby marking
a strict lower limit for the present scenario. Consequently, the RG
improved set-up has a larger domain of validity due to the weakening
of gravity at shorter distances, equally reflected in the boundedness
of the associated Bekenstein-Hawking temperature, see Sec.~\ref{t}.
The second modification takes the quantum gravity-induced reduction of
the event horizon into account, replacing $r_{\rm cl}$ by $r_s$ in
\eq{sc}. This can be written in terms of a form factor, replacing
\eq{sc} by $\hat\sigma=\hat\sigma_{\rm cl}\cdot F(\sqrt{s})$ with
\begin{equation}\label{F}
F(\sqrt{s})=\left.\left(\frac{r_s}{r_{\rm cl}}\right)^2\right|_{M_{\rm phys}=\sqrt{s}}\,,
\end{equation}
see Fig.~\ref{Fs}. We conclude that the RG improved production
cross section is reduced with respect to the semi-classical one,
already in the regime where the semi-classical approximation is
applicable, see Fig.~\ref{Fs}. The quantitative impact of these effects on mini-black
holes production at colliders, eg.~the LHC, is evaluated in \cite{FHL}.
\begin{figure}[t]
\includegraphics[width=.95\hsize]{3DplotF.eps}
\caption{\label{Fs} The gravitational form factor $F(\sqrt{s})$
with parameter $\gamma=\gamma_{\rm dS}$ with $n=4$ extra dimensions.}
\end{figure}
\subsection{Trans-Planckian region}\label{TP}
Next, we implement our RG improvement directly on the level of the
classical Schwarzschild radius rather than on the level of the
underlying black hole metric. To that end, we interpret the energy
dependence of the form factor in the production cross section as
originating from an effective energy dependence of Newton's
coupling. The latter enters the classical event horizon as
\begin{eqnarray} \label{classicalBH}
r_{\rm cl}(\sqrt{s})&=& \frac{1}{\sqrt{\pi}}\left(\frac{8\,\Gamma(\frac{d-1}{2})}{d-2}\right)^{\frac{1}{d-3}}
\left(G_N\,\sqrt{s}\right)^{\frac{1}{d-3}}\quad\quad
\end{eqnarray}
where the substitution $M_{\rm phys}=\sqrt{s}$ has already been
executed. Under the assumption that the functional dependence of
\eq{classicalBH} on the gravitational coupling $G_N$ remains unchanged
once quantum corrections are taken into account, we can interpret the
non-trivial energy-dependence of $r_s(\sqrt{s})$ to originate from
\eq{classicalBH} via the energy-dependence of Newton's
coupling. Substituting $G_N\leftrightarrow G(\sqrt{s})$, we find
\begin{equation}\label{match-s}
{G(\sqrt{s})}={G_N}\,\left[F(\sqrt{s})\right]^\0{d-3}{2}\,.
\end{equation}
Using the non-perturbative form factor $F(\sqrt{s})$, we conclude that
$G(\sqrt{s})$ displays a threshold behaviour, starting off at the
black hole formation threshold $\sqrt{s}\approx M_c$, and increasing
asymptotically towards $G_N$ with increasing
$\sqrt{s}$. Trans-Planckian scattering in gravity becomes classical:
$G(\sqrt{s})$ approaches $G_N$ with increasing center-of-mass energy
$\sqrt{s}\gg M_D$ and the production cross section reduces to the
geometrical one. This is a consequence of quantum corrections being
suppressed for large black hole mass, see Sec.~\ref{pt}, and thus for
large $\sqrt{s}$.
With these results at hand, we can now turn the
argument around and identify the matching $k=k(\sqrt{s})$ which
reproduces \eq{match-s} from the renormalisation group running of
$G(k)$. To leading order in $M_D/\sqrt{s}\ll 1$, the matching
\begin{equation}\label{improved}
G_N\to G(k)\quad{\rm with}\quad k\propto
M_D \left(\frac{M_D}{\sqrt{s}}\right)^{1/(d-3)}
\end{equation}
in \eq{classicalBH} --- together with $G(k)$ from the renormalisation
group, see Sec.~\ref{RG} --- reproduces the form factor \eq{F} and the
semi-classical limit.
The result \eq{improved} highlights a duality between the regime of
large center-of-mass energy $\sqrt{s}/M_D\gg 1$ of a gravitational
scattering process, and the low-momentum behaviour $(k/M_D)^{d-3}
\propto M_D/\sqrt{s}\ll 1$ of the running coupling $G(k)$ in the `gravitational
bound state' of a black hole.
It would be interesting to have access to the behaviour of $G(\sqrt{s})$ at
below-threshold energies, where the energy dependence of Newton's coupling
should be obtained from standard field theory
amplitudes for $s$-channel scattering in asymptotically safe gravity \cite{Litim:2007iu},
which become
strongly dominated
by multi-graviton exchange at Planckian energies \cite{'tHooft:1987rb}. For recent developments
along these lines within quantum string-gravity, see \cite{Amati:2007ak,Marchesini:2008yh}.
\subsection{Semi-classical limit}
It is useful to compare our results with a related renormalisation
group study, where qualitatively different conclusions have been
reached \cite{Koch:2008zzb}. There, black hole production cross
sections are estimated from \eq{classicalBH} using the RG matching
\begin{equation}\label{sk}
G_N\to G(k)\quad{\rm with}\quad k \propto\sqrt{s}
\end{equation}
for Newton's coupling, with $G(k)$ taken from the renormalisation
group and $k$ identified with $\sqrt{s}$, following \cite{Hewett:2007st}.
This would be applicable if
$\sqrt{s}$ is the sole mass scale in the problem, and if $G_N$ in
\eq{classicalBH} is sensitive to the momentum transfer in the
$s$-channel. However, the matching \eq{sk} is in marked contrast to
\eq{improved}. Most importantly, with \eq{sk} no semi-classical limit
is achieved in the trans-Planckian regime, because $G(\sqrt{s})/G_N\ll
1$ becomes strongly suppressed. This conclusion is at variance with
the findings of the present paper.
The origin for this difference is traced back to the following
observation: the RG improved Schwarzschild radius depends on several mass
scales, the Planck scale $M_D$, the black hole mass $M$ and, implicitly,
the momentum scale $k$. Identifying both the mass $M=\sqrt{s}$ and
the renormalisation group scale $k=\sqrt{s}$ with the
center-of-mass energy in a gravitational scattering process entangles
mass dependences with RG scale running. In turn, the d\'etour taken
in Sec.~\ref{TP} disentangles these effects by taking into account
that the physics involves several mass scales. This also
explains why $M_D$ enters the matching \eq{improved}, besides
$\sqrt{s}$, which is responsible for the qualitative difference with
respect to \eq{sk}.
We conclude that the set-up laid out in this work is necessary to capture
the semi-classical limit of trans-Planckian scattering.
\section{Discussion}\label{Discussion}
How does quantum gravity modify the physics of black holes?
We have implemented quantum corrections
on the level of black hole metrics, replacing Newton's constant by
a coupling which runs
under the renormalisation group equations for gravity.
If Newton's coupling weakens
sufficiently fast towards shorter distances, it implies
the existence of a smallest black hole of mass $M_c$.
This is the case for all dimensions $d\ge 4$ provided quantum
gravity is asymptotically safe.
The mass scale $M_c$ is dynamically
generated and of the
order of the fundamental Planck scale $M_D$.
Interestingly, a mere weakening of the gravitational coupling
would not be enough to disallow the formation of an event horizon.
The mechanism responsible for a lower bound on black hole mass
relates with the RG scaling of the gravitational coupling at the
cross-over from perturbative to non-perturbative running.
In consequence, the underlying fixed point is not primarily
responsible for the existence of the lower bound and alternative
UV completions may display a similar weakening down to
length scales of the order of the Planck length.
In the semi-classical regime
$M_D/M\ll 1$, corrections to the event horizon and black hole thermodynamics
remain perturbatively small, but effects become quantitatively more pronounced with
decreasing black hole mass $M$.
Once $M_D/M$ becomes of order one, quantum corrections are more
substantial. The specific heat changes sign, the black hole temperature
displays a maximum, and vanishes with $M\to M_c$.
This supports the view that critical black holes
constitute cold, Planck-size, remnants.
Direct implications of fixed point scaling are visible
in the short distance limit $r M_D\ll 1$. This limit
becomes time-like rather than space-like as in classical Schwarzschild
black holes. Also, asymptotically safe black holes with $M>M_c$
always also display a Cauchy horizon besides the event
horizon. It is noteworthy that the classical curvature singularity at the origin is
significantly softened because of the fixed point, and either
disappears completely, or becomes vastly reduced.
The conformal structure of quantum black holes is
very similar to classical Reissner-Nordstr\"om black holes, including the near
horizon geometry of critical black holes which is of the AdS${}_2\times
{\rm S}^{\rm d-2}$ type.
Our results have direct implications for the collider phenomenology of low-scale
gravity models. Interestingly, quantum corrections increase the domain of validity
for a semiclassical description. At low center-of-mass energies, a threshold
for black hole production is identified. At larger energies, quantum
corrections to production cross sections lead to a new form factor. It reduces
the cross section, and reproduces the semi-classical result
in the trans-Planckian limit. A
quantitative implementation of this scenario for mini-black hole
production is given elsewhere \cite{FHL}.
It would be very interesting to
complement this picture by explicit computations based on multi-graviton
exchange at Planckian energies along the lines laid out in \cite{Litim:2007iu}.\\
{\it Note added.---}
After completion of this work, we have been informed by B.~Koch that some
of our results \cite{erg2008,FHL} have been confirmed in the preprint \cite{Burschil:2009va}.
\bibliographystyle{plain}
|
\section{Introduction}
\label{sec:intro}
In the Internet era, ecommerce has grown and flourished. The greater amount of available information, the lower cost of communication, and other reductions in economic friction makes the world `flatter' than ever before, promoting automated marketplaces and the adoption of autonomous agents in ecommerce \cite{greenwal-jennings-stone-03-iis-agents.and.markets,kephart-02-pnas-software.agents}. In financial markets, traders have continuously turned to automated algorithmic trading services to deal with faster transactions and more complex market dynamics \cite{schartz:francioni:weber:06:equity-trader-course}. According to an article from \textit{The Economist} \cite{070623-economist-algo.trading}, algorithmic trading accounted for a third of all share trades in the United States, and Aite Group, a consultancy, reported that the figure will reach 50\% by 2010. News and events usually affect market predictions and lead to high price volatility, which in turn creates opportunities for arbitrage between markets. Unpredictable dynamics and complex linkages between markets make more robust, efficient market mechanisms very desirable.
Online auction sites like eBay provide a way for consumers to buy a wide range of items. Since its establishment in 1995, eBay has expanded into dozens of countries and now makes billions of dollars each year. The auction mechanisms used by eBay and other successful auction sites however are not perfect. For example, an eBay auction typically finishes at a fixed time, allowing a bidder to bid only moments before the auction terminates and steal a deal from bidders who would offer higher prices if given the chance \cite{greenwald06e-markets}. This means a loss of revenue for both sellers and eBay. Another issue, and one to which many researchers have paid much attention, is that eBay runs many \emph{simultaneous sequential auctions} \cite{gerding-dyj-07-aamas-bidding.in.concurrent.auctions,juda-parkes-06-ec-sequential.ebay.auctions}. In other words, on eBay, hundreds, even thousands, of on-going auctions may sell the same kind of goods. It is difficult for a potential buyer to select an auction that would result in the lowest winning bid. As a result, a successful bid in one auction may be lower than a failed bid in another, leading to complaints from both sellers and bidders, lower efficiency of the auctions and, in time, less revenue for eBay.
Electronic auctions have also been used to sell things that are not goods in a traditional sense. For example, search engines like Google and Yahoo, in the role of publisher, typically use auctions to select and show relevant advertisements along with search results on their web sites. For each keyword-based search query, an \emph{ad auction} is run to select bids from advertisers. Each selected advertiser provides an ad to display in one of a certain number of \emph{ad positions} on the search result page. Better positions, which draw more attention from users, are allotted to advertisers that bid higher. An advertiser usually pays on a \emph{per-click} basis rather than on the \emph{per-impression} basis in the traditional media. Publishers have commonly used variants of an auction mechanism called the \emph{generalized second-priced auction} to determine winning bids from advertisers and their ad positions. Although ad auctions generate many dollars in income each year for these companies, how to analyze the current practices and design more effective ad auction mechanisms is still a major concern. For instance, Lahaie and Pennock \cite{lahaie:pennock:07:revenue-analysis} compared the ranking rule used by Yahoo --- based on the prices of bids --- and that used by Google --- based on the expected profits of bids to Google, and concluded that neither rule consistently outperforms the other.
All these scenarios from e-commerce challenge the designers of electronic auction mechanism to design more desirable mechanisms. This opens up new lines of research in computer science, such as inventing new algorithms for deciding the winning bid in auctions \cite{lehmann-muller-sandholm-06-bookchap-wdp}, deciding how best to bid in multiple auctions \cite{schwartz:byrne:colaninno:08:competition}, and how to build the software infrastructure to run such auctions \cite{niu-cpgmmps-08-aamas-jcat}.
The Internet also significantly boosts the adoption of distributed computing, in particular \emph{agent-based computing}. A major tool for multi-agent system designers has been \emph{game theory} (\acro{gt}). \acro{gt} provides a framework for studying strategic, interacting individuals and solution concepts --- usually various equilibria --- with the assumption of the rationality of individuals. \acro{gt} thus helps to compare the outcomes of an interaction mechanism to the optimal ones in theory, but it does not give a dynamic model that explains how to reach optimal outcomes, nor presents much guidance on how to maximize global outcomes when some agents in the system are not as rational as presumed.
\emph{Auction mechanisms} are an ideal candidate to provide this missing model. The effectiveness of auction mechanisms in the real world and the similarity between an auction and a multi-agent system --- both involving multiple self-interested individuals and concerning certain global outcomes --- have led to various market-based approaches to multi-agent coordination and resource allocation problems in cluster and grid computing environments \cite{horling-lesser-05-ker-organizational.paradigms,stober-neumann-08-ecr-greedex,yeoB06taxonomy}. These approaches have demonstrated superior performance than those pre-existing, non-market solutions, in terms of a combination of performance, scalability, and reliability. However, the market mechanisms adopted by these approaches are usually selected arbitrarily or based on certain heuristics. It is unknown whether these market mechanisms are optimal solutions, or whether there are better options.
\section{A new approach}
\label{sec:strategic-decision-making}
\subsection{Automated mechanism design}
\label{sec:strategic-decision-making:amd}
Facing the challenges in both electronic commerce and market-based control, we need to solve the following problem: \emph{Given a certain set of restrictions and desired outcomes, how can we design a good, if not optimal, auction mechanism; or when the restrictions and goals alter, how can the current mechanism be improved to handle the new scenario?}
The traditional answer to this question has been in the domain of auction theory \cite{krishna-02-book-auction.theory}. A mechanism is designed by hand, analyzed theoretically, and then revised as necessary. The problems with the approach are exactly those that dog any manual process --- it is slow, error-prone, and restricted to just a handful of individuals with the necessary skills and knowledge. In addition, there are classes of commonly used mechanism, such as the double auctions that we discuss here, which are too complex to be analyzed theoretically, at least for interesting cases \cite{walsh-02-gtdt-analyzing.complex.strategic.interactions}.
Automated mechanism design (\acro{amd}) aims to overcome the problems of the manual process by designing auction mechanisms automatically. \acro{amd} considers design to be a search through some space of possible mechanisms. For example, Cliff \cite{cliff-01-tr-evolution.of.market.mech} and Phelps \textit{et al.} \cite{phelps02coev.mechanism.design,phelps03optimizing.pricing.rules} explored the use of evolutionary algorithms to optimize different aspects of the continuous double auction. Around the same time, Conitzer and Sandholm \cite{conitzer-sandholm-03-icec-amd.complexity} were examining the complexity of building a mechanism that fitted a particular specification.
These different approaches were all problematic. The algorithms that Conitzer and Sandholm considered dealt with exhaustive search, and naturally the complexity was exponential. In contrast, the approaches that Cliff and Phelps \textit{et al.} pursued were computationally more appealing, but gave no guarantee of success and were only searching tiny sections of the search space for the mechanisms they considered. As a result, one might consider the work of Cliff and Phelps \textit{et al.}, and indeed the work we describe here, to be what Conitzer and Sandholm \cite{conitzer-sandholm-07-ijcai-incremental} call ``incremental'' mechanism design, where one starts with an existing mechanism and incrementally alters parts of it, aiming to iterate towards an optimal mechanism. Similar work, though work that uses a different approach to searching the space of possible mechanisms has been carried out by \cite{vorobeychik-reeves-wellman-07-uai-constrained.amd} and has been applied to several different mechanism design problems \cite{schvartzman-wellman-09-amec-exploring.large.strategy.spaces}.
The problem with taking the automated approach to mechanism design further is how to make it scale --- though framing it as an incremental process is a good way to look at it, it does not provide much practical guidance about how to proceed. Our aim in this paper is to provide more in the way of practical guidance, showing how it is possible to build on a previous analysis of the most relevant components of a complex mechanism in order to set up an automated mechanism design problem, and then describing one approach to solving this problem.
\subsection{CAT games}
\label{sec:strategic-decision-making:cat}
We set our work within the context of the Trading Agent Competition Market Design game, also known as the \acro{cat} game. This competition, which ran for the last three years, asks entrants to design a market for a set of automated traders which are based on standard algorithms for buying and selling in a double auction, including \acro{zi-c} \cite{gode-sunder-93-jpe-zi}, \acro{zip} \cite{cliff-97-tr-zip}, \acro{re} \cite{erev98predicting}, and \acro{gd} \cite{gjerstad-dickhaut-98-geb-gd}. The game is broken up into a sequence of \emph{days}, and each day every trader picks a market to trade in, using a market selection strategy that models the situation as an $n$-armed bandit problem \cite[Section 2]{sutton-barto-98-book-rl}. Markets are allowed to charge traders in a variety of ways and are scored based on the number of traders they attract (market share), the profits that they make from traders (profit share), and the number of successful transactions they broker relative to the total number of shouts placed in them (transaction success rate). Full details of the game can be found in \cite{cai-gmnpp-09-tr-cat.overview}.
We picked the \acro{cat} game as the basis of our work for four main reasons. First, the double auctions that are the focus of the design are a widely used mechanism. Second, the competition is run using an open-source software package called \acro{jcat} which is a good basis for implementing our ideas. Third, after three years of competition, a number of specialists have been made available by their authors, giving us a library of mechanisms to test against. Fourth, there have been a number of publications that analyze different aspects of previous entrants, giving us a good basis from which to start searching for new mechanisms.
\subsection{Towards a grey-box approach}
\label{sec:strategic-decision-making:white+black}
Particularly helpful is the prior work of Niu \textit{et al.} \cite{niu-cpgm-08-aamas-cats,niu-cmp-08-iat-eco}. \cite{niu-cpgm-08-aamas-cats} is an analysis of the 2007 \acro{cat} competition, which identifies a number of different components of the double auction market, along with different implementations of these components proposed by the game entrants. \cite{niu-cmp-08-iat-eco} complements this analysis with a description of a large number of simulations of competitions between the specialists from the 2007 \acro{cat} game, systematically identifying how different specialists perform both in multi-market games, and in games between pairs of specialists. Together these two papers mirror the black-box and white-box analyses from software engineering. \cite{niu-cpgm-08-aamas-cats} provides a white-box analysis, looking inside each market in order to identify which components it contains, and relating the performance of each market to the operation of its components. \cite{niu-cmp-08-iat-eco} provides a black-box analysis, which ignores the detail of the internal components of each market, but providing a much more extensive analysis of how the specialists perform.
These analyses make a good combination for examining the strengths and weaknesses of specialists. The white-box approach is capable of relating the internal design of a strategy to its performance and revealing which part of the design may cause vulnerabilities, but it requires internal structure and involves manual examination. The black-box approach does not rely upon the accessibility of the internal design of a strategy. It can be applied to virtually any strategic game, and is capable of evaluating a design in many more situations. However, the black-box approach tells us little about what may have caused a strategy to perform poorly and provides little in the way of hints as to how to improve the strategy. It is desirable to combine these two approaches in order to benefit from the advantages of both. Following the \acro{ga}-based approach to trading strategy acquisition and auction mechanism design in \cite{cliff-01-tr-evolution.of.market.mech,phelps06automatic-strategy-acquisition,phelps03optimizing.pricing.rules}, we propose what we call a \emph{grey-box} approach to automated mechanism design that solves the problem of automatically creating a complex mechanism by searching a structured space of auction components.
In other words, we concentrate on the components of the mechanisms (as in the white-box approach), but take a black-box view of the components, evaluating their effectivenesses by looking at their performance against that of their peers.
More specifically, we view a market mechanism as a combination of auction rules, each as an atomic building block. We consider the problem: \emph{how can we find a combination of rules that is better than any known combination according to a certain criterion, based on a pool of existing building blocks?} The black-box analysis in \cite{niu-cmp-08-iat-eco} maintains a population of strategies and evolves them generation by generation based on their fitnesses. Here we intend to follow a similar approach, maintaining a population of components or building blocks for strategies, associating each block with a \emph{quality score}, which reflects the fitnesses of auction mechanisms using this block, exploring the part of the space of auction mechanisms that involves building blocks of higher quality, and keeping the best mechanisms we find.
\section{Grey-box AMD}
\label{sec:greybox}
Having sketched our approach at a high level, we now look in detail at how it can be applied in the context of the \acro{cat} game.
\subsection{A search space of double auctions}
\label{sec:greybox:gda}
The first issues we need to address are \emph{what composite structure is used to represent auction mechanisms?} and \emph{where can we obtain a pool of building blocks?}
Viewing an auction as a structured mechanism is not a new idea. Wurman \textit{et al.} \cite{wurman01parametrization} introduced a conceptual, parameterized view of auction mechanisms. Niu \textit{et al.} \cite{niu-cpgm-08-aamas-cats} extended this framework for auction mechanisms competing in \acro{cat} games and provided a classification of entries in the first \acro{cat} competition that was based on it. The extended framework includes multiple intertwined components, or \emph{policies}, each regulating one aspect of a market. We adopt this framework, include more candidates for each type of policy and take into consideration parameters that are used by these policies.
These policies are either inferred from the literature \cite{mccabe93upda}, or from our previous work \cite{niu-cmp-08-iat-eco,niu-cpgm-08-aamas-cats,niu-cps-06-aamas-pricing.fluctuation}, or contributed by entrants to the \acro{cat} competitions. These policies, each as a building block, form a solid foundation for the grey-box approach.
Figure~\ref{fig:auction-space-tree} illustrates the building blocks as a tree structure which we describe after we review the blocks themselves.
\begin{sidewaysfigure}
\begin{center}
\resizebox{.85\linewidth}{!}{
\includegraphics{figs/tree}
}
\caption{The search space of double auctions modeled as a tree, discussed in details in Section~\ref{sec:greybox}.}
\label{fig:auction-space-tree}
\end{center}
\end{sidewaysfigure}
We describe the different types of policies in details below and discuss how we search the space based on the tree structure in the next section.
\subsubsection{Matching policy}
\label{sec:greybox:gda:matching}
Matching policies, denoted as \ar{M} in Figure~\ref{fig:auction-space-tree}, define how a market matches shouts made by traders.
\emph{Equilibrium matching} (\ar{me}) is the most commonly used matching policy \cite{mccabe93upda,wurman-98-dss-flexible.das}. The offers made by traders form the \emph{reported demand and supply}, which is usually different from the \emph{underlying demand and supply}, and are determined by traders' private values and unknown to the market, since traders are assumed to be profit-seeking and make offers deviating from their private values. \ar{me} clears the market at the \emph{reported} equilibrium price and matches intra-marginal asks (offers to sell) with intra-marginal bids (offers to buy) --- with an intersecting demand and supply, the shouts on the left of the intersection (the equilibrium point) and their traders are called \emph{intra-marginal} since they can be matched and make profit, while those on the right are called \emph{extra-marginal}. It is worth mentioning that a shout, or a trader, that appears to be intra-marginal or extra-marginal in the reported demand and supply may not be so in the underlying demand and supply.
\emph{Max-volume matching} (\ar{mv}) aims to increase transaction volume based on the observation that a high intra-marginal bid can match with a lower extra-marginal ask, though with a profit loss for the buyer. It does so to realize the maximal transaction volume that is possible.
A generic, parameterized, matching policy can be defined to include \ar{me} and \ar{mv} as two special cases. This policy, denoted as \ar{mt}, uses a parameter, $\theta$, which can be any value in $[-1,1]$. When $\theta$ is $-1$, \ar{mt} does not match any shout; when $\theta$ is $0$, \ar{mt} becomes \ar{me}; and when $\theta$ is $1$, \ar{mt} becomes \ar{mv}. For any other values of $\theta$, \ar{mt} tries to realize a transaction volume that is proportional to $0$ and those realized in \ar{me} and \ar{mv}.
\subsubsection{Quote policy}
\label{sec:greybox:gda:quoting}
Quote policies, denoted as \ar{Q} in Figure~\ref{fig:auction-space-tree}, determine the quotes issued by markets. Typical quotes are ask and bid quotes, which respectively specify the upper bound for asks and the lower bound for bids that may be placed in a \emph{quote-driven} market. \emph{Two-sided quoting}\fn{The name follows \cite{mccabe93upda} since either quote depends on information on both the ask side and the bid side.} (\ar{qt}) defines the ask quote as the minimum of the lowest tentatively matchable bid and lowest unmatchable ask, and defines the bid quote as the the maximum of the highest tentatively matchable ask and highest unmatchable bid.
\emph{One-sided quoting} (\ar{qo}) is similar to \ar{qt}, but considers only the standing shouts closest to the reported equilibrium price from the unmatchable side. When the market is cleared continuously (see below), \ar{qo} is identical to \ar{qt}, but otherwise forms a possibly looser restriction on placing shouts.
\emph{Spread-based quoting} (\ar{qs}) extends \ar{qt} to maintain a higher ask quote and a lower bid quote for use with \ar{mv}. With \ar{qs}, when the ask quote is lower than the bid quote, the former is set somewhere above their average and the latter below the average, and the spread between the two is a fixed value. \ar{qs} helps relax the constraint put on shouts with too low an ask quote and too high a bid quote.
\subsubsection{Shout accepting policy}
\label{sec:greybox:gda:accepting}
Shout accepting policies, denoted as \ar{A} in Figure~\ref{fig:auction-space-tree}, judge whether a shout made by a trader should be permitted in the market.
\emph{Always accepting} (\ar{aa}) accepts any shout, and \emph{never accepting} (\ar{an}) does the opposite.
\emph{Quote-beating accepting} (\ar{aq}) allows only those shouts that are more competitive than the corresponding market quote. This has been commonly used in both experimental settings and real stock markets, and is sometimes called the ``New York Stock Exchange rule'' since that market adopts it.
\emph{Self-beating accepting} (\ar{as}) accepts all first-time shouts but only allows a trader to modify its standing shout with a more competitive price.
\emph{Equilibrium-beating accepting} (\ar{ae}) learns an estimate of the equilibrium price based on the past transaction prices in a sliding window, and requires bids to be higher than the estimate and asks to be lower. \ar{ae} uses a parameter, $w$, to specify the size of the sliding window in terms of the number of transactions, and a second parameter, $\delta$, which can be added to the estimate to relax the restriction on shouts. This policy was suggested in \cite{niu-cps-06-aamas-pricing.fluctuation} and found to be effective in reducing transaction price fluctuation and increasing allocative efficiency in markets populated by \acro{zi-c} traders.
A variant of \ar{ae}, denoted as \ar{ad} and introduced by the \s{PSUCAT} team in the first \acro{cat} competition, uses the standard deviation of transaction prices in the sliding window rather than a constant $\delta$ to relax the restriction on shouts.
\emph{History-based accepting} (\ar{ah}) is derived from the \acro{gd} trading strategy \cite{gjerstad-dickhaut-98-geb-gd} and reported to be a crucial component of one particular strong market mechanism for \acro{cat} games \cite{niu-cmp-08-iat-eco}. \acro{gd} computes how likely a given offer is to be matched, based on the history of previous shouts, and \ar{ah} uses this to accept only shouts that will be matched with probability no lower than a specified threshold, $\tau\in [0,1]$.
\emph{Transaction-based accepting} (\ar{at}) tracks the most recently matched asks and bids, and uses the lowest matched bid and the highest matched ask to restrict the shouts to be accepted. In a clearing house (\acro{ch}) \cite{friedman-rust-93-book-dam}, the two bounds are expected to be close to the estimate of equilibrium price in \ar{ae}, while in a continuous double auction (\acro{cda}), \ar{at} may produce much looser restriction since extra-marginal shouts may steal a deal.
\emph{Shout type-based accepting} (\ar{ay}) allows shouts based merely on their types, i.e. asks or bids. This mimics the continuum of auctions presented in \cite{cliff-01-tr-evolution.of.market.mech}, including retailer markets where only sellers shout, procurement auctions where only buyers shout, as well as general double auctions.
\subsubsection{Clearing condition}
\label{sec:greybox:gda:clearing}
Clearing conditions, denoted as \ar{C} in Figure~\ref{fig:auction-space-tree}, define when to clear the market and execute transactions between matched asks and bids.
\emph{Continuous clearing} (\ar{cc}) attempts to clear the market whenever a new shout is placed.
\emph{Round clearing} (\ar{cr}) clears the market after all traders have submitted their shouts. This was the original clearing policy in \acro{nyse}, but was replaced by \ar{cc} later for faster transactions and higher volumes. With \ar{cc}, an extra-marginal trader may have more chance to steal a deal and get matched.
\emph{Probabilistic clearing} (\ar{cp}) clears the market with a predefined probability, $p$, whenever a shout is placed. It thus defines a continuum of clearing rules with \ar{cr} ($p=0$) and \ar{cc} ($p=1$) being the two ends.
\subsubsection{Pricing policy}
\label{sec:greybox:gda:pricing}
Pricing policies, denoted as \ar{P} in Figure~\ref{fig:auction-space-tree}, set transaction prices for matched ask-bid pairs. The decision making may involve only the prices of the matched ask and bid, or more information including market quotes.
\emph{Discriminatory $k$-pricing} (\ar{pd}) sets the transaction price of a matched ask-bid pair at some point in the interval between their prices. The parameter $k\in [0,1]$ controls which point is used and usually takes value $0.5$ to avoid a bias in favor of buyers or sellers.
\emph{Uniform $k$-pricing} (\ar{pu}) is similar to \ar{pd}, but sets the transaction prices for all matched ask-bid pairs at same point between the ask quote and the bid quote. A transaction price set by \ar{pu} may or may not fall into the range between the matched ask and bid, depending upon the matching policy and the quote policy in the auction mechanism, When it falls outside, whichever of the ask and the bid is closer to the computed transaction price will be used as the final transaction price.
\emph{$n$-pricing} (\ar{pn}) was introduced in \cite{niu-cps-06-aamas-pricing.fluctuation}, and sets the transaction price as the average of the latest $n$ pairs of matched asks and bids. If the average falls out of the price interval between the ask and bid to be matched, the nearest end of the interval is used. This policy can help reduce transaction price fluctuation and has little impact on allocative efficiency.
\emph{Side-biased pricing} (\ar{pb}) is basically \ar{pd} with an internal $k$ dynamically adjusted so as to split the profit in favor of the side on which fewer shouts exist. Thus the more that asks outnumber bids in the current market, the closer $k$ is set to 0.
\subsubsection{Charging policy}
\label{sec:greybox:gda:charging}
Charging policies, denoted as \ar{G} in Figure~\ref{fig:auction-space-tree}, determine the charges imposed by a market. This is typically not an issue in research on auctions in isolation, but would affect the selection of markets by traders directly in an environment of multiple competing markets, as in \acro{cat} games.
\emph{Fixed charging} (\ar{gf}) sets charges at a specified fixed level.
\emph{Bait-and-switch charging} (\ar{gb}) makes a market cut its charges until it captures a certain market share, and then slowly increases charges to increase profit. It will adjust its charges downward again if its market share drops below a certain level.
\emph{Charge-cutting charging} (\ar{gc}) sets the charges by scaling down the lowest charges imposed on the previous day, based on the observation that traders prefer markets with lower charges.
\emph{Learn-or-lure-fast charging} (\ar{gl}) adapts charges towards some target following the scheme used by the \acro{zip} trading strategy \cite{cliff-97-tr-zip}. If the market using this policy believes that the traders are still exploring among markets and have yet to find a good one to trade, the market would adapt charges towards 0 to lure traders to join and stay; otherwise it learns from the charges of the most profitable market. \ar{gl} uses an exploring monitor component to determine whether traders are exploring or not. A simple exploring monitor, for example, examines the daily distribution of market shares of specialists. If the distribution is flat, the traders are considered exploring, and not otherwise. This is based on the observation that traders all tend to go to the best market and cause an imbalanced distribution. To implement this, the degree of flatness of the distribution --- the standard deviation of the distribution relative to the mean of the distribution --- is compared to a threshold $\tau \in [0,1]$. And similar to \acro{zip}, \ar{gl} uses a learning rate parameter, $r$, to control how fast the market adapts its charges.
All these charging policies require an initial set of fees on different activities, including fee on registration, fee on information, fee on shout, fee on transaction, and fee on profit, denoted as $f_r$, $f_i$, $f_s$, $f_t$, and $f_p$ respectively in Figure~\ref{fig:auction-space-tree}.
\subsubsection{A tree model}
\label{sec:greybox:gda:tree}
The tree model of double auctions in Figure~\ref{fig:auction-space-tree} illustrates how building blocks are selected and assembled level by level. There are \emph{and} nodes, \emph{or} nodes, and \emph{leaf} nodes in the tree. An \emph{and} node, rounded and filled, combines a set of building blocks, each represented by one of its child nodes, to form a compound building block. The root node, for example, is an \emph{and} node to assemble policies, one on each aspect described above, to obtain a complete auction mechanism. An \emph{or} node, rectangular and filled, represents the decision making of selecting a building block from the candidates represented by the child nodes of the \emph{or} node based on their quality scores. This selection occurs not only for those major aspects of an auction mechanism, i.e. \ar{m}, \ar{q}, \ar{a}, \ar{p}, \ar{c}, and \ar{g} (at \ar{g}'s child node of `policy' in fact), but also for minor components, for example, a learning component for an adaptive policy (as what Phelps \textit{et al.} does regarding a trading strategy \cite{phelps06automatic-strategy-acquisition}), and for determining optimal values of parameters in a policy, like $\theta$ in \ar{mt} and $k$ in \ar{pd}. A \emph{leaf} node represents an atomic block that can either be for selection at its \emph{or} parent node or be further assembled into a bigger block by its \emph{and} parent node. A special type of \emph{leaf} node in Figure~\ref{fig:auction-space-tree} is that with a label in the format of $[x,y]$. Such a \emph{leaf} node is a convenient representation of a set of \emph{leaf} nodes that have a common parent --- the parent of this special \emph{leaf} node --- and take values evenly distributed between $x$ and $y$ for the parameter labeled at the parent node.
\emph{or} nodes contribute to the variety of auction mechanisms in the search space and are where exploitation and exploration occur. We model each \emph{or} node as an $n$-armed bandit learner that chooses among candidate blocks, and use the simple softmax method \cite[Section 2.3]{sutton-barto-98-book-rl} to solve this learning problem. The same solution is adopted in designing the market selection strategy for trading agents in \acro{cat} games \cite{niu-cps-07-tada-market.competition}.\footnote{However the two scenarios may need different parameter values. The market selection scenario should favor choices that give a good profit --- a cumulative measure --- while here we require effective exploration to find a good mechanism in the foreseeable future --- a one-time concern.}
\subsection{The \textsc{Grey-Box-AMD} algorithm}
\label{sec:greybox:algorithm}
\newcommand{\ms{EM}}{\ms{EM}}
\newcommand{\ms{HOF}}{\ms{HOF}}
\newcommand{\ms{FM}}{\ms{FM}}
\newcommand{\ms{SM}}{\ms{SM}}
\newcommand{\ms{M}}{\ms{M}}
\renewcommand{\S}{\ms{S}}
\newcommand{\ms{G}}{\ms{G}}
\newcommand{\ms{X}}{\ms{X}}
\newcommand{\ms{Y}}{\ms{Y}}
\newcommand{\const}[1]{\texttt{\textsc{#1}}}
\newcommand{\kw}[1]{\textbf{#1}}
Given a set of building blocks, $\ms{B}$, and a set of fixed markets, $\ms{FM}$, as targets to beat, we define the skeleton of the grey-box algorithm below:
\medskip
\begin{algorithm}{Grey-Box-AMD}{\ms{B}, \ms{FM}}
\ms{HOF} \= \{\} \\
\begin{FOR}{s \= 1 \TO \const{num\_of\_steps}}
G \= \CALL{Create-Game}() \\
\ms{SM} \= \{\} \\
\begin{FOR}{m \= 1 \TO \const{num\_of\_samples}}
M \= \CALL{Create-Market}() \\
\begin{FOR}{t \= 1 \TO \const{num\_of\_policytypes}}
B \= \CALL{Select}(\ms{B}_t, 1) \\
\CALL{Add-Block}(M, B)
\end{FOR} \\
\ms{SM} \= \ms{SM} \cup \{M\}
\end{FOR} \\
\ms{EM} \= \CALL{Select}(\ms{HOF}, \const{num\_of\_hof\_samples})) \\
\CALL{Run-Game}(G, \ms{FM}\cup\ms{EM}\cup\ms{SM}) \\
\begin{FOR}{\EACH M \IN \ms{EM} \cup \ms{SM}}
\CALL{Update-Market-Score}(M, \CALL{Score}(G, M)) \\
\begin{IF}{M\ \kw{not} \IN \ms{HOF}}
\ms{HOF} \= \ms{HOF} \cup \{M\}
\end{IF} \\
\begin{IF}{\const{capacity\_of\_hof} < |\ms{HOF}|}
\ms{HOF} \= \ms{HOF} - \{\CALL{Worst-Market}(\ms{HOF})\}
\end{IF} \\
\begin{FOR}{\EACH B \text{used by} M}
\CALL{Update-Block-Score}(B, \CALL{Score}(G, M))
\end{FOR}
\end{FOR}
\end{FOR} \\
\RETURN \ms{HOF}
\end{algorithm}
The \textsc{Grey-Box-AMD} algorithm runs a certain number of steps (\const{num\_of\_steps} in Line 2). At each step, a single \acro{cat} game is created (\textsc{Create-Game}() in Line 3) and a set of markets are prepared for the game. This set of markets includes all markets in $\ms{FM}$, a certain number (\const{num\_of\_samples} in Line 5) of markets sampled from the search space, denoted as $\ms{SM}$, and a certain number (\const{num\_\-of\_\-hof\_samples} in Line 11) of markets, denoted as $\ms{EM}$, chosen from a Hall of Fame, $\ms{HOF}$. All these markets are put into the game, which is run to evaluate the performance of these markets (\textsc{Run-Game}($G$, $\ms{FM}\cup\ms{EM}\cup\ms{SM}$) in Line 12). $\ms{HOF}$ has a fixed capacity, \const{capacity\_of\_hof}, and maintains markets that performed well in games at previous steps in terms of their average scores across games they participated. $\ms{HOF}$ is empty initially, updated after each game, and returned in the end as the result of the grey-box process.
Each market in $\ms{SM}$ is constructed based on the tree model in Figure~\ref{fig:auction-space-tree}. After an `empty' market mechanism, $M$, is created (\textsc{Create-Market}() in Line 6), building blocks can be incorporated into it (\textsc{Add-Block}($M$,$B$) in Line 9, where $B\in\ms{B}$). \const{num\_of\_policytypes} in Line 7 defines the number of different policy types, and from each group of policies of same type, denoted as $\ms{B}_t$ where $t$ specifies the type, a building block is chosen for $M$ (\textsc{Select}($\ms{B}_t$, $1$) in Line 8). For simplicity, this algorithm illustrates only what happens to the \emph{or} nodes at the high level, including \ar{M}, \ar{Q}, \ar{A}, \ar{C}, and \ar{P}. Markets in $\ms{EM}$ are chosen from $\ms{HOF}$ in a similar way (\textsc{Select}($\ms{HOF}$, \const{num\_of\_hof\_samples}) in Line 11).
After a \acro{cat} game, $G$, completes at each step, the game score of each participating market $M\in\ms{SM}\cup\ms{EM}$, \textsc{Score}($G$, $M$), is recorded and the game-independent score of $M$, \textsc{Score}($M$), is updated (\textsc{Update-Market-Score}($M$, \textsc{Score}($G$, $M$)) in Line 14). If $M$ is not currently in $\ms{HOF}$ and \textsc{Score}($M$) is higher than the lowest score of markets in $\ms{HOF}$, it replaces that corresponding market (\textsc{Worst-Market}($\ms{HOF}$) in Line 18).
\textsc{Score}($G$, $M$) is also used to update the quality score of each building block used by $M$ (\textsc{Update-Block-Score}($B$, \textsc{Score}($G$, $M$)) in Line 20). Both \textsc{Update-Market-Score} and \textsc{Update-Block-Score} calculate respectively game-independent scores of markets and quality scores of building blocks by averaging feedback \textsc{Score}($G$, $M$) over time. Because choosing building blocks occurs only at \emph{or} nodes in the tree, only child nodes of an \emph{or} node have quality scores and receive feedback after a \acro{cat} game. Initially, quality scores of building blocks are all $0$, so that the probabilities to choose them are even. As the exploration proceeds, fitter blocks score higher and are chosen more often to construct better mechanisms.
\section{Experiments}
\label{sec:experiments}
This section describes the experiments that are carried out to acquire auction mechanisms using the grey-box approach.
\subsection{Experimental setup}
\label{sec:experiments:setup}
We extended \acro{jcat} with the parameterized framework of double auctions and all the individual policies described in Section~\ref{sec:greybox:gda}. To reduce the computational cost, we eliminated the exploration of charging policies by focusing on mechanisms that impose a fixed charge of $10\%$ on trader profit, which we denote as $\ar{gf}_{0.1}$. Analysis of \acro{cat} games \cite{niu-cmp-08-iat-eco} and what entries have typically charged in actual \acro{cat} competitions, especially in the latest two events, suggest that such a charging policy can be a reasonable choice to avoid losing either intra-marginal or extra-marginal traders. Even with this cut-off, the search space still contains more than $1,200,000$ different kinds of auction mechanisms, due to the variety of policies on aspects other than charging and the choices of values for parameters.
The experiments that we ran to search the space each last 200 steps. At each step, we sample two auction mechanisms from the space, and run a \acro{cat} game to evaluate them against four fixed, well known, mechanisms plus two mechanisms that performed well at previous steps and are members of the Hall of Fame. The scores of the sampled and Hall of Fame mechanisms are used as feedback for every building block that an individual mechanism uses and is associated with a quality score.
To sample auction mechanisms, the softmax exploration method used by \emph{or} nodes starts with a relatively high temperature so as to explore randomly, then gradually cools down, and eventually maintains a temperature that guarantees a non-negligible probability of choosing even the worst action any time. After all, our goal in the grey-box approach is not to converge quickly to a small set of mechanisms, but to explore the space as broadly as possible and avoid being trapped in local optima.
The fixed set of four markets in every \acro{cat} game includes two \acro{ch} markets --- $\acro{ch}_l$ and $\acro{ch}_h$ --- and two \acro{cda} markets --- $\acro{cda}_l$ and $\acro{cda}_h$ --- with one of each charging $10\%$ on trader profit, like $\ar{gf}_{0.1}$ does, and the other charging $100\%$ on trader profit (denoted as $\ar{gf}_{1.0}$). \acro{ch} and \acro{cda} mechanisms are two common double auctions and have been used in the real world for many years, in financial marketplaces in particular due to their high allocative efficiency. Earlier experiments we ran, involving \acro{ch} and \acro{cda} markets against entries into \acro{cat} competitions, indicate that it is not trivial to win over these two standard double auctions. Markets with different charge levels are included to avoid any sampled mechanisms taking advantage otherwise. Based on the parameterized framework in Section~\ref{sec:greybox:gda}, the \acro{ch} and \acro{cda} markets can be represented as follows:
\begin{center}
\begin{tabular}{r@{ / }l@{ = }l}
$\acro{ch}_l$ & $\acro{ch}_h$ & \ar{me} + \ar{qt} + \ar{aq} + \ar{cr} + $\ar{pu}_{k=0.5}$ + $\ar{gf}_{0.1}$ / $\ar{gf}_{1.0}$ \\
$\acro{cda}_l$ & $\acro{cda}_h$ & \ar{me} + \ar{qt} + \ar{aq} + \ar{cc} + $\ar{pd}_{k=0.5}$ + $\ar{gf}_{0.1}$ / $\ar{gf}_{1.0}$
\end{tabular}
\end{center}
The Hall of Fame that we maintain during the search contains ten `active' members and a list of `inactive' members. After each \acro{cat} game, the two sampled mechanisms are compared with those active Hall of Famers. If the score of a sampled mechanism is higher than the lowest average score of the active Hall of Famers, the sampled mechanism is inducted into the Hall of Fame and replaces the corresponding Hall of Famer, which becomes inactive and ineligible for \acro{cat} games at later steps. An inactive Hall of Famer may be reactivated if an identical mechanism happens to be sampled from the space again and scores high enough to promote its average score to surpass the lowest score of active Hall of Famers.
Each \acro{cat} game is populated by 120 trading agents, using \acro{zi-c}, \acro{zip}, \acro{re}, and \acro{gd} strategies, a quarter of the traders using each strategy. Half the traders are buyers, half are sellers. The supply and demand schedules are both drawn from a uniform distribution between 50 and 150. Each \acro{cat} game lasts 500 days with ten rounds for each day. This setup is similar to that of actual \acro{cat} competitions except for a smaller trader population that helps to reduce computational costs. A 200-step grey-box experiment takes around sixteen hours on a \acro{windows pc} that runs at 2.8GHz and has a 3GB memory.
\subsection{Experimental results}
\label{sec:experiments:results}
We carried out four experiments to check whether the grey-box approach is successful in searching for good auction mechanisms.
First, we measured the performance of the mechanisms that are being generated indirectly, through their effect on other mechanisms. Since the four standard markets participate in all the \acro{cat} games, their performance over time reflects the strength of their opponents --- they will do worse as their opponents get better --- which in turn reflects whether the search generates increasingly better mechanisms. Figure~\ref{fig:fixed-markets} shows that the scores of the four markets (more specifically the average daily scores of the markets in a game) decrease over 200 games, especially over the first 100 games, suggesting that the mechanisms we are creating get better as the learning process progresses.
Second, we measured the performance of the set of mechanisms we are creating more directly. The mechanisms that are active in the Hall of Fame at a given point represent the best mechanisms that we know about at that point and their performance tells us more directly how the best mechanisms evolve over time. Figure~\ref{fig:elite} shows the scores of the ten active Hall of Famers at each step over a 200-step run.\fn{Note that the active Hall of Famers will be different mechanisms at different steps in the process, so what we see in the figure is the performance of the best mechanisms we know of.} As in Figure~\ref{fig:fixed-markets}, the first 100 steps sees a clear, increasing trend. Note that even the scores of the worst of the ten at the end are above 0.35, higher than the highest of the four fixed markets from Figure~\ref{fig:fixed-markets}. Thus we know that our approach will create mechanisms that outperform standard mechanisms, though we should not read too much into this since we trained our new mechanisms directly against them.
\begin{figure}[tb]
\begin{center}
\subfloat[The four fixed auction mechanisms.]{\label{fig:fixed-markets}\includegraphics[scale=0.4]{figs/fixed-markets}}
\qquad
\subfloat[The top ten active Hall of Famers.]{\label{fig:elite}\includegraphics[scale=0.4]{figs/elites}}
\caption{Scores of market mechanisms over 200 steps during the grey-box process.}
\label{fig:scores}
\end{center}
\end{figure}
Third, a better test of the new mechanisms is to run them against those mechanisms that we know to be strong in the context of \acro{cat} games, asking what would have happened if our Hall of Fame members had been entered into prior \acro{cat} competitions and had run against the carefully hand-coded entries in those competitions. We chose three Hall of Famers, which are internally labeled as \s{SM7.1}, \s{SM88.0}, and \s{SM127.1} and can be represented in the parameterized framework in Section~\ref{sec:greybox:gda} as follows:
\begin{center}
\begin{tabular}{r@{ = }l}
\s{SM7.1} & \ar{mv} + \ar{qo} + $\ar{ah}_{\tau=0.4}$ + $\ar{cp}_{p=0.3}$ + $\ar{pn}_{n=11}$ + $\ar{gf}_{0.1}$ \\
\s{SM88.0} & $\ar{mt}_{\theta=0.4}$ + \ar{qt} + \ar{aa} + $\ar{cp}_{p=0.4}$ + $\ar{pu}_{k=0.7}$ + $\ar{gf}_{0.1}$ \\
\s{SM127.1} & \ar{mv} + \ar{qs} + \ar{as} + $\ar{cp}_{p=0.4}$ + $\ar{pu}_{k=0.7}$ + $\ar{gf}_{0.1}$ \\
\end{tabular}
\end{center}
We ran these three mechanisms against the best recreation of past \acro{cat} competitions that we could achieve given the contents of the \acro{tac} agent repository,\fn{\url{http://www.sics.se/tac/showagents.php}.} where competitors are asked to upload their entries after the competition. There were enough entries in the repository at the time we ran the experiments to create reasonable facsimiles of the 2007 and 2008 competitions, but there were not enough entries from the 2009 competition for us to recreate that year's competition. The \acro{cat} games were set up in a similar way to the competitions, populated by 500 traders that are evenly split between buyers and sellers and between the four trading strategies --- \acro{zi-c}, \acro{zip}, \acro{re}, and \acro{gd} --- and the private values of sellers or buyers were drawn from a uniform distribution between 50 and 150. For each recreated competition, we ran three games, like in the actual competitions.
Table~\ref{tab:catscores} lists the average cumulative scores of all the markets across their three games along with the standard deviations of those scores. The three new mechanisms we obtained from the grey-box approach beat the actual entries in the competition by a comfortable margin in both cases. The fact that we can take mechanisms that we generate in one series of games (against the fixed opponents and other new entries) and have them perform well against a separate set of mechanisms suggests that the grey-box approach learns robust mechanisms.
In passing, we note that the rankings of the entries from the repository do not reflect those in the actual \acro{cat} competitions. This is to be expected since the entries now face much stronger opponents and different markets will, in general, respond differently to this. Excluding the markets that attempt to impose invalid fees and are marked with `*', we can see that the overall performance of entries into the 2008 \acro{cat} competition is better than that of those into the 2007 \acro{cat} competition when they face the three new, strong, opponents, reflecting the improvement in the entries over time.
\begin{table*}[tb]
\begin{center}
\caption{The scores of markets in \acro{cat} games including the best mechanisms from the grey-box approach and entries in prior \acro{cat} competitions, averaged over three \acro{cat} games respectively.}
\label{tab:catscores}
\subfloat[Against \acro{cat} 2007 entries.]{\label{tab:catscores-07}
\begin{tabular}{+lrr}
\toprule
\textbf{Market} & \textbf{Score} & \textbf{\acro{sd}}\\
\midrule[\heavyrulewidth]
\s{SM7.1} & 199.4500 & 5.9715 \\
\s{SM88.0} & 191.1083 & 10.3186 \\
\s{SM127.1} & 180.1277 & 9.0289 \\
\s{MANX} & 154.6953 & 1.3252 \\
\s{CrocodileAgent} & 142.0523 & 9.0867 \\
\s{TacTex} & 138.4527 & 5.8224 \\
\s{PSUCAT} & 133.1347 & 5.6565 \\
\s{PersianCat} & 124.3767 & 11.2409 \\
\s{jackaroo} & 108.8017 & 8.6851 \\
\s{IAMwildCAT}\tmark[*] & 106.8897 & 4.4006 \\
\s{Mertacor} & 89.1707 & 4.9269 \\
\ \\
\bottomrule
\end{tabular}
}
\qquad
\qquad
\qquad
\subfloat[Against \acro{cat} 2008 entries.]{\label{tab:catscores-08}
\begin{tabular}{+lrr}
\toprule
\textbf{Market} & \textbf{Score} & \textbf{\acro{sd}}\\
\midrule[\heavyrulewidth]
\s{SM7.1} & 196.7240 & 9.2843 \\
\s{SM88.0} & 186.9247 & 4.2184 \\
\s{SM127.1} & 183.5887 & 9.7835 \\
\s{jackaroo} & 177.5913 & 2.5722 \\
\s{Mertacor} & 161.5440 & 5.8741 \\
\s{MANX} & 147.3050 & 15.7718 \\
\s{IAMwildCAT} & 142.9167 & 8.9581 \\
\s{PersianCat} & 139.1553 & 17.9783 \\
\s{DOG} & 130.2197 & 18.9782 \\
\s{MyFuzzy} & 125.9630 & 1.9221 \\
\s{CrocodileAgent}\tmark[*] & 71.4820 & 5.8687 \\
\s{PSUCAT}\tmark[*] & 68.3143 & 6.7389 \\
\bottomrule
\end{tabular}
}
\begin{minipage}{0.99\linewidth}
{\scriptsize{*}} \s{IAMwildCAT} from \acro{cat} 2007, and \s{CrocodileAgent} and \s{PSUCAT} from \acro{cat} 2008 worked abnormally during the games and tried to impose invalid fees, probably due to competitions from the three new, strong opponents. Although we modified \acro{jcat} to avoid kicking out these markets on those trading days when they impose invalid fees --- which \acro{jcat} does in an actual \acro{cat} tournament --- these markets still perform poorly, in contrast to their rankings in the tournaments.
\end{minipage}
\end{center}
\end{table*}
Finally, we tested the performance of \s{SM7.1}, \s{SM88.0}, and \s{SM127.1} when they are run in isolation, applying the same kind of test that auction mechanisms are traditionally subject to.
We tested the mechanisms both for allocative efficiency and, following \cite{niu-cps-06-aamas-pricing.fluctuation}, for the extent to which they trade close to theoretical equilibrium as measured by the coefficient of convergence, $\alpha$. Niu \textit{et al.} \cite{niu-cps-06-aamas-pricing.fluctuation} compared a class of double auctions, called \acro{ncdaee}, which can be represented as:
\begin{center}
\acro{ncdaee} = \ar{me} + $\ar{ae}_{w,\delta}$ + \ar{cc} + $\ar{pn}_n$
\end{center}
The advantage of \acro{ncdaee} is that it can give significantly lower $\alpha$ --- faster convergence of transaction prices --- and higher allocative efficiency ($E_a$) than a \acro{cda} when populated respectively by homogeneous \acro{zi-c} traders and can perform comparably to a \acro{cda} when populated by homogeneous \acro{gd} traders.
We replicated these experiments using \acro{jcat} and ran additional ones for the three new mechanisms with similar configurations. The results of these experiments are shown in Table~\ref{tab:economic}.\fn{Our results are slightly different from those in \cite{niu-cps-06-aamas-pricing.fluctuation}, but the pattern of these results still holds. In addition, we ran an \acro{ncdaee} variant ($\delta=30$) that was not tested in \cite{niu-cps-06-aamas-pricing.fluctuation}, observing that those with $\delta\leq 20$ do not perform well when populated by \acro{gd} traders.} The best result in each column is shaded. We can see that both cases of \s{SM7.1} with \acro{zi-c} traders and \s{SM88.0} with \acro{gd} traders give higher $E_a$ than the best of the existing markets respectively, and both of these increases are statistically significant at the 95\% level. Both cases also lead to low $\alpha$, not the lowest in the column but close to the lowest, and the differences between them and the lowest are not statistically significant at the 95\% level. Thus the grey-box approach can generate mechanisms that perform as well in the single market scenario as the best mechanisms from the literature.
\begin{table*}[tb]
\begin{center}
\caption{Economic properties of the best mechanisms from the grey-box experiments and the auction mechanisms explored in \protect\cite{niu-cps-06-aamas-pricing.fluctuation}. All \acro{ncdaee} mechanisms are configured to have $w=4$ in their \ar{ae} policies and $n=4$ in their \ar{pn} policies.The best result in each column is shaded. Data in the first four row are averaged over 1,000 runs and those in the last four are averaged over 100 runs.}
\label{tab:economic}
\begin{tabular}{+lrrrrrrrr}
\toprule\rowstyle{\centering}
\multirow{3}*{\textbf{Market}} & \multicolumn{4}{c}{\textbf{\acro{zi-c}}} & \multicolumn{4}{c}{\textbf{\acro{gd}}} \\ \cmidrule (l){2-5} \cmidrule (l){6-9}
& \multicolumn{2}{c}{\textbf{$E_a$}} & \multicolumn{2}{c}{\textbf{$\alpha$}} & \multicolumn{2}{c}{\textbf{$E_a$}} & \multicolumn{2}{c}{\textbf{$\alpha$}} \\ \cmidrule (l){2-3} \cmidrule (l){4-5} \cmidrule (l){6-7} \cmidrule (l){8-9}
& \textbf{Mean} & \textbf{\acro{sd}} & \textbf{Mean} & \textbf{\acro{sd}} & \textbf{Mean} & \textbf{\acro{sd}} & \textbf{Mean} & \textbf{\acro{sd}} \\
\midrule[\heavyrulewidth]
\acro{cda} & 97.464 & 3.510 & 13.376 & 4.351 & 99.740 & 1.553 & \cellcolor[gray]{0.9}4.360 & 3.589 \\
$\acro{ncdaee}_{\delta=0}$ & 98.336 & 3.262 & \cellcolor[gray]{0.9}4.219 & 3.141 & 9.756 & 28.873 & 14.098 & 1.800 \\
$\acro{ncdaee}_{\delta=10}$ & 98.912 & 2.605 & 5.552 & 2.770 & 23.344 & 41.727 & 7.834 & 5.648 \\
$\acro{ncdaee}_{\delta=20}$ & 98.304 & 2.562 & 7.460 & 3.136 & 89.128 & 30.867 & 4.826 & 3.487 \\
$\acro{ncdaee}_{\delta=30}$ & 97.708 & 3.136 & 8.660 & 3.740 & 99.736 & 1.723 & 4.498 & 3.502 \\ \midrule
\s{SM7.1} & \cellcolor[gray]{0.9}99.280 & 1.537 & 4.325 & 2.509 & 58.480 & 47.983 & 4.655 & 4.383 \\
\s{SM88.0} & 98.320 & 2.477 & 11.007 & 4.251 & \cellcolor[gray]{0.9}99.920 & 0.560 & 4.387 & 2.913 \\
\s{SM127.1} & 97.960 & 3.225 & 11.152 & 4.584 & 99.520 & 1.727 & 4.751 & 3.153 \\
\bottomrule
\end{tabular}
\end{center}
\end{table*}
\section{Conclusions and future work}
\label{sec:conclusions}
This paper describes a practical approach to the automated design of complex mechanisms. The approach that we propose breaks a mechanism down into a set of components each of which can be implemented in a number of different ways, some of which are also parameterized. Given a method to evaluate candidate mechanisms, the approach then uses machine learning to explore the space of possible mechanisms, each composed from a specific choice of components and parameters. The key difference between our approach and previous approaches to this task is that the score from the evaluation is not only used to grade the candidate mechanisms, but also the components and parameters, and new mechanisms are generated in a way that is biased towards components and parameters with high scores.
The specific case-study that we used to develop our approach is the design of new double auction mechanisms. Evaluating the candidate mechanisms using the infrastructure of the TAC Market Design competition, we showed that we could learn mechanisms that can outperform the standard mechanisms against which learning took place and the best entries in past Market Design competitions. We also showed that the best mechanisms we learnt could outperform mechanisms from the literature even when the evaluation did not take place in the context of the Market Design game. These results make us confident that we can generate robust double auction mechanisms and, as a consequence, that the grey-box approach is an effective approach to automated mechanism design.
Now that we can learn mechanisms effectively, we plan to adapt the approach to also learn trading strategies, allowing us to co-evolve mechanisms and the traders that operate within them.
\section*{Acknowledgments}
This work is partially supported by the \acro{US NSF} under grant IIS-0329037, \emph{Tools and
Techniques for Automated Mechanism Design}, and the \acro{UK EPSRC} under grants GR/T10657/01 and GR/T10671/0, \emph{Market Based Control of Complex Computational Systems}. We thank the entrants to the Market Design game for releasing binaries of their market agents.
|
Subsets and Splits